development of new routes for the synthesis of

266
Illinois State University Illinois State University ISU ReD: Research and eData ISU ReD: Research and eData Theses and Dissertations 5-3-2017 Development Of New Routes For The Synthesis Of Development Of New Routes For The Synthesis Of Carbaporphyrins And Carbachlorins Carbaporphyrins And Carbachlorins Navneet Sahota Illinois State University, [email protected] Follow this and additional works at: https://ir.library.illinoisstate.edu/etd Part of the Chemistry Commons Recommended Citation Recommended Citation Sahota, Navneet, "Development Of New Routes For The Synthesis Of Carbaporphyrins And Carbachlorins" (2017). Theses and Dissertations. 773. https://ir.library.illinoisstate.edu/etd/773 This Thesis is brought to you for free and open access by ISU ReD: Research and eData. It has been accepted for inclusion in Theses and Dissertations by an authorized administrator of ISU ReD: Research and eData. For more information, please contact [email protected].

Upload: others

Post on 18-Dec-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Illinois State University Illinois State University

ISU ReD: Research and eData ISU ReD: Research and eData

Theses and Dissertations

5-3-2017

Development Of New Routes For The Synthesis Of Development Of New Routes For The Synthesis Of

Carbaporphyrins And Carbachlorins Carbaporphyrins And Carbachlorins

Navneet Sahota Illinois State University, [email protected]

Follow this and additional works at: https://ir.library.illinoisstate.edu/etd

Part of the Chemistry Commons

Recommended Citation Recommended Citation Sahota, Navneet, "Development Of New Routes For The Synthesis Of Carbaporphyrins And Carbachlorins" (2017). Theses and Dissertations. 773. https://ir.library.illinoisstate.edu/etd/773

This Thesis is brought to you for free and open access by ISU ReD: Research and eData. It has been accepted for inclusion in Theses and Dissertations by an authorized administrator of ISU ReD: Research and eData. For more information, please contact [email protected].

DEVELOPMENT OF NEW ROUTES FOR THE SYNTHESIS OF CARBAPORPHYRINS

AND CARBACHLORINS

Navneet Sahota

250 Pages

Carbaporphyrins and carbachlorins are porphyrin analogs in which one of the nitrogens is

replaced with a carbon atom within a porphyrin-like cavity. These systems form stable

organometallic derivatives and exhibit unusual chemical reactivity. Most of the studies in this area

have been carried out with benzocarbaporphyrins, where one of the pyrrole rings is replaced by an

indene unit. Although syntheses of monocarbaporphyrins with fused benzene rings have been

widely investigated over the past 20 years, few examples of carbaporphyrins without fused rings

have been reported in the literature.

In this thesis, two projects were conducted to further explore carbaporphyrin and

carbachlorin chemistry. The first project utilized a ‘3+1’ condensation of a tripyrrane dicarboxylic

acid with dimethyl 3,5-diformylcyclopentane-1,2-dicarboxylate in the presence of trifluoroacetic

acid, followed by oxidation with 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ), to yield a

novel carbaporphyrin system in 40% yield together with a 9% yield of a related carbachlorin. This

new carbaporphyrinoid system retained macrocyclic aromaticity and afforded stable silver(III) and

gold(III) complexes. In addition, alkylation with methyl iodide and potassium carbonate gave a

C-methyl derivative, and this underwent metalation with palladium(II) acetate and nickel(II)

acetate to afford related organometallic species. The second project dealt with the development

of a new route for synthesizing carbachlorins. Chlorins are a group of dihydroporphyrins and are

important biological pigments. The study of core-modified chlorins has been limited compared to

that of core modified porphyrins. Carbachlorins are chlorin analogues where carbon atom is

inserted into the chlorin core. They have interesting properties and are also promising candidates

in many practical applications. Based on a “3+1” methodology, a novel carbachlorin was

synthesized and characterized through various spectroscopic techniques, including NMR, UV-Vis

spectroscopy, and mass spectrometry. The coordination chemistry of this carbachlorin was also

investigated.

KEYWORDS: Carbaporphyrins, Carbachlorins, Porphyrins, Chlorins, Spectroscopic techniques

DEVELOPMENT OF NEW ROUTES FOR THE SYNTHESIS OF CARBAPORPHYRINS

AND CARBACHLORINS

NAVNEET SAHOTA

A Thesis Submitted in Partial

Fulfillment of the Requirements

for the Degree of

MASTER OF SCIENCE

Department of Chemistry

ILLINOIS STATE UNIVERSITY

2017

Copyright 2017 Navneet Sahota

DEVELOPMENT OF NEW ROUTES FOR THE SYNTHESIS OF CARBAPORPHYRINS

AND CARBACHLORINS

NAVNEET SAHOTA

COMMITTEE MEMBERS:

Timothy D. Lash, Chair

Shawn R. Hitchcock

Christopher C. Mulligan

i

ACKNOWLEDGMENTS

The author wishes to acknowledge all of the members of her research group for all of their

support throughout the duration of her time at Illinois State University. The author would also like

to acknowledge the faculty, staff, and fellow students in the Chemistry Department at Illinois State

University. The writer acknowledges financial support from the National Science Foundation

under grants no. CHE-1465049 and CHE-1212691 and the Petroleum Research Fund.

The author would like to thanks the members of her committee, Dr. Hitchcock and Dr.

Mulligan for all the guidance given. In addition, the author acknowledges Dr. Ferrence for

providing X-ray crystal structures for a carbaporphyrin and related metalloporphyrinoids. Great

thanks are given to her research advisor Dr. Timothy D. Lash for his encouragement, endless

support, and patience. The writer feels truly privileged and thankful to have had such an excellent

mentor for the past two years.

N.S

ii

CONTENTS

Page

ACKNOWLEDGMENTS i

CONTENTS ii

FIGURES iv

SCHEMES vii

CHAPTER

I. INTRODUCTION TO PORPHYRINS 1

Introduction 1

Structure of Porphyrins 3

Aromatic Properties 7

Biosynthesis of Heme 16

Synthesis of Porphyrins 19

Reactivity of Porphyrins 30

Macrocycles Related to Porphyrins 32

Applications 34

Carbaporphyrins 36

Carbachlorins 50

II. SYNTHESIS OF CARBAPORPHYRINS 66

Introduction 66

Results and Discussion 67

Conclusions 109

Experimental 111

III. SYNTHESIS OF A CARBACHLORIN 126

Introduction 126

Results and Discussion 129

Conclusions 141

Experimental 142

IV. SUMMARY AND CONCLUSIONS 150

REFERENCES 152

APPENDIX A: SELECTED NMR SPECTRA 156

iii

APPENDIX B: SELECTED UV-VIS SPECTRA 229

iv

FIGURES

Figure Page

1. Resonance Structures of Pyrrole 4

2. Examples of Cyclic Structures Depicting

Differences in Aromaticity 8

3. Structures of Porphyrin and [18]Annulene 9

4. Reduced Forms of Porphyrin 11

5. Induced Ring Current Effects on Porphyrin as Observed in

Nuclear Magnetic Resonance Spectroscopy 13

6. UV-Vis Spectrum of Tetraphenylporphyrin 14

7. Structures of Dipyrrolic Intermediates 23

8. Structures of Tripyrrane Intermediates Used in ‘3+1’ Condensations 29

9. Carbaporphyrins Stabilizing Metals with Rare Oxidation States 31

10. Structures of Contracted and Expanded Porphyrins 33

11. Structures of Phthalocyanine and Dicarbahemiporphyrazine 33

12. Monocarbaporphyrin Analogues 37

13. Mono-, Di-, Tri- and Tetracarbaporphyrins 38

14. Aromatic Delocalization Pathways in Porphyrin and Chlorin 51

15. Structures of Chlorin and Carbachlorin 53

16. 500 MHz 1H NMR Spectrum of Fumarate Diol in CDCl3 77

17. 500 MHz Proton NMR Spectrum of Carbaporphyrin 98 in CDCl3 80

18. 500 MHz Proton NMR Spectrum of the Carbachlorin byproduct in CDCl3 81

19. UV-Vis Spectrum of Carbaporphyrin 98 in Dichloromethane 82

20. POV-Ray generated ORTEP III Drawing of 98 83

v

21. UV-Vis Spectrum of 117 in Dichloromethane 84

22. 500 MHz Proton NMR Spectrum of Mixture of a 98H+ and 98H22+ in CDCl3 85

23. UV-Vis Spectrum of Carbaporphyrin 98H+ in CH2Cl2 86

24. 500 MHz Proton NMR Spectrum of 98H+ in CDCl3 87

25. 500 MHz 1H NMR Spectrum of 98H22+ in CDCl3 88

26. 500 MHz Proton NMR Spectra of 98 Showing Exchange of 21-H

after 10 min Exposure to d-TFA 91

27. 500 MHz proton NMR Spectra of 98 Showing Partial

Exchange of the meso-protons after 24 hr Exposure to d-TFA 91

28. 500 MHz 1H NMR Spectrum of 117H� in CDCl3 93

29. UV-Vis Spectrum of 117H+ By-product in 5% TFA-Dichloromethane. 94

30. 500 MHz 1H NMR Spectrum of 118 in CDCl3 95

31. 500 MHz 1H NMR Spectrum of 120 in CDCl3 98

32. 500 MHz 1H NMR Spectrum of 121 in CDCl3 100

33. 500 MHz 1H NMR Spectrum of 120H� in CDCl3 101

34. 500 MHz 1H NMR Spectrum of 120H22+ in CDCl3 102

35. 500 MHz 1H NMR Spectrum of 121H+ in CDCl3 103

36. 500 MHz 1H NMR Spectrum of 122 in CDCl3 105

37. UV-Vis Spectrum of Palladium(II) Carbaporphyrin in Chloroform 105

38. POV-Ray rendered ORTEP III Drawing of 122 106

39. 500 MHz 1H NMR Spectrum of 123 in CDCl3 108

40. POV-Ray rendered ORTEP III Drawing of 123 108

41. 500 MHz Proton NMR Spectrum of Carbachlorin 133 in CDCl3 133

42. UV-Vis Spectrum of Carbachlorin 133 134

vi

43. 500 MHz Proton NMR Spectrum of Carbachlorin 133H+ in CDCl3 135

44. Proton NMR Spectrum of Carbachlorin 133 in CDCl3 with Two Drops of d-TFA

after 20 min (upper spectrum) and 7 hours (lower spectrum) 138

45. 500 MHz Proton NMR Spectrum of 138 in CDCl3 140

46. UV-Vis Spectrum of Silver(III) Chlorin 138 in Dichloromethane 141

vii

SCHEMES

Scheme Page

1. Mechanism of the Knorr Pyrrole Synthesis 5

2. Mechanism for the Hantzsch Synthesis of Substituted Pyrroles 6

3. Tautomerization and Protonation of Porphyrin 7

4. Synthesis of Uroporphyrinogen III 17

5. Biosynthesis of Heme from Uroporphyrinogen III 18

6. Rothemund Synthesis of TPP 20

7. Adler-Longo synthesis of TPP 21

8. Synthesis of meso-substituted porphyrins by Lindsey’s method 22

9. Fischer Porphyrin Synthesis 23

10. MacDonald's “2+2” condensation reaction 24

11. Synthesis of Porphyrin from an Oxophlorin 25

12. Synthesis of Oxaporphyrins and Thiaporphyrins 26

13. Synthesis of the First Tripyrrane 27

14. Synthesis of Tripyrranes Reported by Sessler and Coworkers 28

15. Porphyrin Synthesis through a Tetrapyrrolic Intermediate 30

16. Bromination of Porphyrins 32

17. Synthesis of Benzocarbaporphyrin 39

18. Synthesis of Tetraaryl Benzocarbaporphyrins

from Tetraarylazuliporphyrins 40

19. Protonation of Benzocarbaporphyrin 41

20. Synthesis of Silver(III) Benzocarbaporphyrin 42

21. Synthesis of Gold(III) Benzocarbaporphyrin 42

viii

22. Formation of Ir and Rh complexes of Benzocarbaporphyrin 44

23. Alkylation of Benzocarbaporphyrin 45

24. Unexpected Alkyl Group Migration 45

25. Oxidation of Benzocarbaporphyrin with FeCl3 48

26. Synthesis of Dibenzo opp-Dicarbaporphyrin 49

27. Protonation of Dibenzo opp-Dicarbaporphyrin 50

28. Phlorin-Chlorin Conversion 52

29. Synthesis of Carbachlorins 54

30. Synthesis of Carbaporphyrins and Attempted Synthesis of a Carbachlorin 55

31. Retrosynthesis of Desired Dialdehyde 56

32. Synthesis of Carbachlorin 57

33. Protonation of Carbachlorin 58

34. Oxidation of Carbachlorin 59

35. Formation of Silver(I) Complex of Carbachlorin 61

36. Formation of N-methylated Carbachlorin 62

37. Unexpected Formation of Pd Complex 64

38. Retrosynthetic Analysis of Carbaporphyrin 98 67

39. Retrosynthetic Analysis of the Tripyrrane Intermediate 68

40. Preparation of Pyrrole Ester 101 using a Knorr-Type Synthesis 69

41. Formation of a Pyrrole Benzyl Ester 69

42. Formation of an Acetoxymethylpyrrole 70

43. Formation of Ethyl Isocyanoacetate 105 70

44. Preparation of 3,4-Diethylpyrrole via a Barton-Zard Synthesis 72

ix

45. Preparation of Tripyrrane Dicarboxylic Acid 62 73

46. Preparation of Bicyclic Adduct 111 73

47. Formation of Diol 112 74

48. Unsuccessful Attempts to Synthesize Diol 112 75

49. Formation of Bicyclic Adduct 114 76

50. Formation of Diol 115 76

51. Formation of Dialdehyde 99 77

52. Mechanism for Ring Opening 78

53. Preparation of Carbaporphyrin 98 79

54. Protonation of Carbaporphyrin 98 86

55. Protonation of Carbaporphyrin 98 88

56. Formation of Different Cationic Species in Equilibrium 90

57. Protonation of Carbachlorin byproduct 117H+ 92

58. Synthesis of Silver(III) Carbaporphyrin 118 95

59. Formation of Gold(III) Complex of Carbaporphyrin 96

60. Formation of C-methylated Carbaporphyrin 120 97

61. Formation of Dimethylated Carbaporphyrin 121 99

62. Metalation of C-methylcarbaporphyrin with Palladium(II) Acetate 104

63. Metalation of C-methylcarbaporphyrin with Nickel(II) Acetate 107

64. Reduction of Carbaporphyrin 98 110

65. New Class of Carbaporphyrin Derivatives 111

66. Retrosynthetic Analysis of Carbachlorin 133 129

67. Formation of 136 130

x

68. Formation of 135 130

69. Formation of Dialdehyde 134 131

70. Preparation of Carbachlorin 133 132

71. Protonation of 133 135

72. Proposed Equilibrium of Different Cationic Species for Carbachlorin 133 137

73. Formation of Silver(III) Chlorin 138 139

1

CHAPTER I

INTRODUCTION TO PORPHYRINS

Introduction

Porphyrins and porphyrinoids are important heterocyclic macrocycles that are widespread

in nature.1 They are usually deeply colored green, red or purple fluorescent crystalline pigments

whose origins can be either synthetic or natural.2 The parent porphyrin system is called porphin 1,

while substituted porphins are called porphyrins. The word porphyrin was derived from the Greek

word porphyra meaning purple. It was in 1873 that a German chemist named Hoppe-Seyler

isolated a purple crystalline compound from heating hemoglobin with sulfuric acid, and he

subsequently coined the name hematoporphyrin for this material from the Greek words for blood

and purple.2 Over the last 150 years, porphyrins have been widely investigated. The interesting

properties of porphyrins, which include their coordination chemistry and aromatic character, have

stimulated research into this class of organic compounds. Porphyrins and their derivatives greatly

contribute to the beauty of nature and play important roles in biological systems. For these reasons,

they have been commonly referred to as the pigments or colors of life.2

2

As mentioned above, many porphyrins and porphyrin-like compounds play significant

roles in both animals and plants. In mammals, heme (2), an iron complex of porphyrin, is a

component of hemoglobin and myoglobin. The former is involved in oxygen transportation around

the body in the blood, while the latter stores oxygen in muscle tissue. Heme is also involved in

redox reactions and electron transfer processes.2 For instance, if the external environment of the

heme is changed, it can facilitate completely different iron chemistry. Thus, in cytochrome c, it

performs the function of electron transfer in cell respiration, as the iron atom cycles through +2

and +3 oxidation states.

In vitamin B12, nature makes use of a coordinated cobalt ion within a corrin core 3. The

complex is used to transfer hydrogen atoms and to reduce organic species. Vitamin B12 is a water-

soluble vitamin and plays essential roles in the formation of red blood cells and in maintaining

healthy nerve cells. Lack of vitamin B12 leads to a disease called pernicious anemia where the

shortage of red cells and hemoglobin leads to impairment of the central nervous system. This

disease usually results from the imperfect absorption of the vitamin by the intestines.2

In plants, chlorophylls (4), magnesium complexes of porphyrin-like structures, capture

photons of light in the near-ultraviolet and red regions of the visible spectrum which therefore

make them important for photosynthesis. The magnesium ion in the reduced porphyrin macrocycle

of chlorophyll serves to modulate the energy transfer and light absorbing characteristics of

chlorophyll, while acting as a center for binding water. A number of chlorophyll molecules are

arranged into tiny molecular antennae that harvest large quantities of photons.1 The energy of the

photons is trapped as excited electrons, which are led away by a chain of electron-transporting

proteins to generate high energy molecules necessary for turning carbon dioxide into

carbohydrates during photosynthesis.2

3

Structure of Porphyrins

The macrocyclic structure of porphyrin was first proposed by Küster in 1912.1 The

macrocycle is composed of four pyrrole units connected by methine bridges at the α-carbons.1,2

The methine carbons are also referred as meso-carbons and the hydrogens attached to these carbons

are known as meso-protons. The pyrrole unit is a five-membered aromatic ring consisting of four

carbon atoms and one nitrogen atom and possesses two formal double bonds. Pyrrole is an

electron-rich aromatic system and has four dipolar resonance contributors (Figure 1) due to the

presence of an electron-donating nitrogen atom.3

4

Pyrrole is essentially non-basic since the lone pair of electrons on the nitrogen is

delocalized into the ring in order to make the structure aromatic. Due to the resonance structures

of this molecule, pyrrole is relatively electron-rich and therefore is likely to undergo electrophilic

aromatic substitution.3 The chemical and physical properties of porphyrins are very different from

the properties of the pyrrole. For instance, porphyrins are usually very deeply colored compounds

whereas pyrroles are off-white solids or colorless liquids. In addition, two of the nitrogens within

the porphyrin core are basic and the π-electrons are delocalized over the entire macrocycle.2

A large number of synthetic routes have been developed for preparing pyrroles. Among

these, the two of the most common methods used to synthesize substituted pyrroles are the Knorr

pyrrole synthesis and the Hantzsch pyrrole synthesis.2 In the Knorr pyrrole synthesis (Scheme 1),

a substituted pyrrole 5 is formed by the condensation of an α-amino ketone 6 with a β-dicarbonyl

compound 7. In order to form the reactive α-amino ketone, an oxime 8 is reduced using zinc dust

in acetic acid.2

Figure 1 Resonance Structures of Pyrrole

5

The second most common method is the Hantzsch pyrrole synthesis. In this approach, an

α-haloketone is reacted with a β-ketoester and a primary amine or ammonia.2 Nucleophilic attack

by the amine or ammonia occurs on the carbonyl carbon of a β-ketoester 9, and subsequent

elimination of water results in the formation of an enamine 10. Condensation of enamine 10 with

the α-haloketone 11, followed by ring closure, yields the substituted pyrrole 12 (Scheme 2).2

Scheme 1 Mechanism of the Knorr Pyrrole Synthesis

6

Scheme 2 Mechanism for the Hantzsch Synthesis of Substituted Pyrroles

Porphyrins can exist in several tautomeric forms, three of which are shown below

(structures 1a, 1b and 1c). Although structures 1a and 1b are equivalent and are interconverting

tautomers, tautomer 1c is less favored due to the close proximity of two adjacent internal

hydrogens that results in steric crowding effects.2 In the porphyrin core, there are two pyrrolenine

nitrogens, which can undergo protonation in the presence of acid to form mono- (1H+) or dications

(1H22+) (Scheme 3).

7

Scheme 3 Tautomerization and Protonation of Porphyrin

Aromatic Properties

Porphyrins are highly conjugated compounds, and it is the conjugated nature of these

macrocycles that is, in part, responsible for their aromatic characteristics.1,2 The concept of

aromaticity was first introduced by Kekulé in 1865.4 The term aromatic was proposed by

Hoffmann because many benzene derivatives have a sweet odor.5 As studies of aromatic

compounds progressed, it was seen that a common property for these compounds was not their

aroma but their relatively high stability. In 1931, Hückel was able to mathematically predict the

aromaticity of unsaturated cyclic systems.2 This led to the well-known Hückel’s rule for

aromaticity, according to which there needs to be [4n+2]π electrons for a molecule to be aromatic,

where n is zero or any positive integer.

Aromatic compounds are found to be more stable when compared to unsaturated

compounds that are not aromatic, and the stability is due to the delocalized electron cloud also

8

known as resonance energy. Along with having sufficient π electrons, it is also required that the

molecule is sufficiently planar, cyclic and continuously conjugated. The more planar the molecule,

the easier it is for the π electrons to interact with one another to create a delocalized system.2

Benzene is the classic system which illustrates aromaticity. It is cyclic, has 6π electrons

and is continuously conjugated, so it follows Hückel’s rule for aromaticity and is very stable. Like

benzene, porphyrin also follows Hückel’s rule for aromaticity. Although porphyrins contain 22

conjugated π electrons (26π electrons counting the lone pairs of electrons on the nitrogen atoms),

only 18 of them participate in a continuous delocalization pathway and therefore follow Hückel’s

rule where n = 4.2 Unsaturated systems can also be classified as anti-aromatic or non-aromatic

(Figure 2). Hückel described anti-aromatic species as cyclic and having a fully conjugated [4n]π

electron pathway. Anti-aromatic compounds are still planar, but are considerably less stable than

aromatic compounds (e.g. cyclobutadiene). On the other hand, non-aromatic structures are

unsaturated systems that are not fully conjugated and may be non-planar. For instance,

cyclooctatetraene is non-aromatic and adopts a tub-like shape making the structure non-planar.

Non-aromatic polyenes are quite reactive as they are relatively unstable.3

Figure 2 Examples of Cyclic Structures Depicting

Differences in Aromaticity

9

Benzene and related compounds are unusually unreactive and do not have the typical

reactivity that is associated with olefins. For example, in the case of cyclooctatetraene, which is

an unusually reactive non-aromatic species, Br2 can readily add across the double bonds giving

addition products. However, benzene is quiet stable and unreactive, and no addition reactions are

observed.3 In benzene or other aromatic compounds, due to the interaction of the π electrons to

create a delocalized π electron cloud, the bond lengths are found to be in between those of single

bonds and double bonds.

Porphyrins can be considered to be the [18]annulenes of nature.6 Annulenes are completely

conjugated hydrocarbon macrocycles with the general formula of CnHn (where n is an even

number) or CnHn+1 (where n is an odd number), n here stands for the number of carbon atoms in

the ring. Annulenes, like other cyclic systems, can be aromatic, non-aromatic or anti-aromatic.

[18]Annulene (13) is an aromatic compound with nine double bonds. The macrocycle is

reasonably planar and has 18π electrons that are fully delocalized over 18 carbon atoms. Therefore

it follows Hückel’s [4n+2]π electron rule for aromaticity, where n = 4. [18]Annulene was first

synthesized by Sondheimer et al. in 19626 (Figure 3), and its conjugation pathway closely

resembles the aromatic circuit present in porphyrins.

Figure 3 Structures of Porphyrin and [18]Annulene

10

Reduced porphyrin systems may also retain aromatic properties. Chlorins 14

bacteriochlorins 15 and isobacteriochlorins 16 all possess 18 π electron delocalization pathways

and are aromatic compounds. Porphyrins can be reduced under a variety of conditions to generate

hydroporphyrins such as chlorins, in which two or more hydrogens have been added across the β-

β double bonds of one or more pyrrole moieties to produce pyrroline rings 14 (Figure 4). Chlorins

are relatively stable compounds that can be oxidized back up to the porphyrin by high potential

quinones such as 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ) and chloranil. In chlorins,

reduction at the β-pyrrole carbons does not interrupt the continuously conjugated delocalisation

pathway. In fact, two more hydrogens may be added across another β-β double bond to give tetra-

hydroporphyrins while retaining the 18 π electron pathway. If the double bonds are reduced on

opposite pyrrole units bacterochlorins are formed, but reduction of the adjacent pyrrole units

affords isobacterochlorin.2

However, other reduced forms of porphyrin, such as phlorins 17, may not be aromatic like

chlorins. This is due to phlorin containing a reduced double bond at one of the bridges linking two

pyrrole units together.2 Without the methine bridge, continuous conjugation throughout the

macrocycle is not possible, therefore breaking one of Hückel’s rules for aromaticity.

11

Figure 4 Reduced Forms of Porphyrin

A number of factors can lead to distortions from planarity in the porphyrin macrocycle. For

instance, the presence of bulky substituents on the macrocyclic system can inhibit the planarity

due to steric interactions. In addition, distortions can arise from the binding of a metal ion within

the cavity of the macrocyclic system. When an appropriate sized metal cation binds in the cavity

of the porphyrin, it may increase the planarity of the molecule. However, if the binding metal

cation is too small or too big for the cavity, the porphyrin ring will distort itself in order to bind to

the metal and hence the planarity will decrease making the ring system potentially less aromatic.

For example, 2,3,7,8,12,13,17,18-octaethylporphyin (OEP) in its free base form is quite planar.

However, when OEP binds to nickel(II), which is a relatively small cation for the cavity of the

porphyrin, the macrocycle twists out of a planar conformation in order to coordinate with the metal.

Porphyrins can tolerate considerable distortion while retaining aromatic properties.2

The aromatic character of porphyrins can be readily assessed using proton nuclear

12

magnetic resonance (NMR) spectroscopy and UV-vis spectroscopy. In particular, the diatropic

character of these systems can easily be analyzed using proton NMR spectroscopy. When these

macrocycles are introduced into a magnetic field, their π electrons circulate to create an aromatic

ring current. The concept of aromatic ring currents was first introduced by Pauling in relation to

the anisotropic diamagnetic susceptibilities of benzenoid hydrocarbons.7 These susceptibilities

were explained in terms of an induced circulation of the delocalized π-electrons in the presence of

an applied magnetic field that produces an induced magnetic field. The strength of the induced

magnetic field generated by the ring current is dependent on the orientation of the aromatic system

with respect to the applied magnetic field. Typically, for porphyrins, the diamagnetic ring current

deshields the β-pyrrole protons and the meso-protons. Because the meso-protons are attached to

electron-deficient carbons, they are shifted even further downfield than the β-pyrrole protons and

appear near +10 ppm. However, the inner N-H protons are shifted far upfield, beyond

tetramethylsilane (TMS, the arbitrary standard for NMR measurements), showing up around -4

ppm. This is because the induced ring current, which deshields protons from the external magnetic

field outside the macrocycle, shields them if they are inside the macrocycle (Figure 5). This

shielding effect is not observed in the case of benzene as there are no internal protons. However,

in some larger aromatic molecules (e.g. [18]annulene), the internal protons may be observed

diamagnetically shielded and hence shifted upfield.

13

Figure 5 Induced Ring Current Effects on Porphyrin

as Observed in Nuclear Magnetic Spectroscopy

The diamagnetic ring currents observed in the case of porphyrins are valuable indicators

of variations in aromatic character. Upon the addition of acid, porphyrins are mono-protonated or

diprotonated, and the resulting cationic species show even stronger diamagnetic ring currents. For

example, in the proton NMR spectrum of porphyrin dications, the meso-protons are shifted slightly

further downfield, while the internal NH protons move upfield.7

Porphyrins have their own characteristic UV-visible spectra. Like benzene, and its

homologues, there are mainly two distinct regions in the porphyrin absorption spectra. However,

unlike benzene, porphyrins are deeply colored compounds due to the absorptions appearing in the

near-ultraviolet and visible regions of the electromagnetic spectrum.1 Highly conjugated systems

tend to absorb light in the near UV and visible regions of their electronic spectra because

conjugated π bonds are able to absorb light at higher wavelengths when compared to isolated π

bonds. Thus, increasing the π bond conjugation of a system increases the wavelengths for

absorption in the UV-Vis spectrum, and this leads to them being colored compounds.

Porphyrins characteristically exhibit intense peaks around 400 nm in the near ultraviolet

region.2 In 1883, Soret,8 while studying hemoglobin, noticed these intense peaks and they were

14

later named as “Soret bands” or B bands. At higher wavelengths, other less intense peaks known

as Q bands are also observed. The intensity of the Soret bands has been correlated to the

aromaticity of the system. The more aromatic the macrocycle is, the more intense the Soret band.2

In the case of tetraphenylporphyrin, an intense Soret band is observed at 417 nm and four Q bands

can be seen at 514 nm, 549 nm, 589 nm, and 646 nm (Figure 6).7

Figure 6 UV-Vis Spectrum of Tetraphenylporphyrin

15

Chlorins 14, although they have less π electron density, still exhibit a Soret band in their

UV-Vis spectra because these structures retain an 18π electron pathway. However, phlorins 17 do

not exhibit Soret bands as there is an interruption of the 18π electron conjugated system due to

reduction at one of the methine bridges. Similarly, porphyrinogens, the reduced intermediates in

the biosynthesis of porphyrins, are also not conjugated structures and in this case are colorless

compounds.2

The differences in the colors of porphyrinoid macrocycles are due to variations in their

absorptions in the visible spectrum. For example, unlike porphyrins, chlorins tend to have a

characteristic absorption near 650 nm, and this results in their characteristic green coloration. This

explains why chlorophyll, a magnesium chlorin, is excellent at absorbing red light.2

The number and intensity of the Q and B bands can give powerful clues about the

substitution pattern of porphyrins and whether they are metalated or not. Protonation of the

porphyrin also alters the UV-vis spectrum. Diprotonation of the macrocycle leads to increased

symmetry compared to the free base form, and this produces a simplification of the Q band region.

Essentially, the four Q bands collapse to two and this can lead to profound color changes.

Substitution at the β-pyrrole positions can also affect the UV-visible absorption spectra

which in turn leads to color changes. Metalloporphyrins tend to be more symmetrical macrocycles

than free-base porphyrins and for this reason they generally only have two Q bands.1 The relative

intensities of these two bands can be used as a qualitative check to see how well the metal cation

fits into the macrocyclic cavity.

16

Biosynthesis of Heme

As mentioned above, in 1912 Küster was the first to propose the structure of porphyrin.9

However, the existence of such a large macrocyclic structure was only fully accepted in 1929,

when Fischer synthesized protoheme.10 The biosynthesis of porphyrins takes place through several

enzyme-catalyzed steps, starting with 5-aminolevulinic acid (ALA). The biosynthesis of

hemoglobin, chlorophyll and vitamin B12 all begin with this precursor. The two substrates required

for the biosynthesis of ALA are succinyl coenzyme A (CoA) and the amino acid glycine. Succinyl

CoA is produced during the Krebs cycle (also known as the citric acid cycle).

In the first step of the pathway, succinyl CoA is condensed with glycine, catalyzed by ALA

synthase, to generate ALA. The condensation of two molecules of ALA, with the aid of ALA

dehydratase, forms the pyrrole intermediate porphobilinogen (PBG). Four pyrrole units then

condense together, in the third step mediated by PBG deaminase, to form the open chain

tetrapyrrolic structure hydroxymethylbilane. Then, uroporphyrinogen III cosynthetase facilitates

the cyclization of the tetrapyrrolic intermediate to form uroporphyrinogen III (Scheme 4). The

cyclization here is associated with an inversion of ring D to afford the type III arrangement of

substituents that can be observed in nearly all naturally occurring porphyrin derivatives. The

biosynthesis of vitamin B12 also branches from the uroporphyrinogen III precursor.2

17

Scheme 4 Synthesis of Uroporphyrinogen III

The enzyme uroporphyrinogen decarboxylase mediates the decarboxylation of all four

acetic acid side chains, while coproporphyrinogen oxidase subsequently oxidatively

decarboxylates only two of the propionic acid side chains. Oxidation of protoporphyrinogen IX by

the enzyme protoporphyrinogen IX oxidase then yields protoporphyrin IX. Insertion of iron(II)

into protoporphyrin IX gives the metalloporphyrin heme (Scheme 5). Heme is then inserted into

the appropriate protein to give hemoglobin or myoglobin, which are both vital moieties in many

biological processes. Protoporphyrin IX is also the biosynthetic precursor to the chlorophylls.2

18

Scheme 5 Biosynthesis of Heme from Uroporphyrinogen III

19

Synthesis of Porphyrins

Over the years, a number of routes have been developed for porphyrin synthesis. Some of

the earliest methods for porphyrin synthesis involved the degradation of naturally occurring

pigments such as heme and chlorophyll. Metalloporphyrins are abundant in nature. For example,

heme derived from blood or chlorophyll from plants, are convenient starting materials that can be

used to prepare a wide variety of porphyrins.2 Although, porphyrins can be synthesized from

different natural sources, this approach can only be used to prepare structures with similar

substitution patterns. For this reason, multistep syntheses have also been developed. The synthetic

methods for the preparation of porphyrins and analogous systems can involve the intermediacy of

monopyrrolic, dipyrrolic, tripyrrolic or tetrapyrrolic structures.2,11

Porphyrins can easily be formed by the tetramerization of monopyrroles. This technique is

mainly used for the synthesis of highly symmetrical porphyrins such as tetraphenylporphyrin or

octaethylporphyrin, and more sophisticated methods are usually required to prepare

unsymmetrical structures.11

In the 1930s, tetraphenylporphyrin was first synthesized by Rothemund, albeit in low

yields.12 He initially investigated the synthesis of porphyrins by reacting acetaldehyde with pyrrole

in methanol at different temperatures in sealed vessels. Based on the same strategy, he extended

the scope of this methodology by reacting pyrrole 18 with several other aldehydes, including

formaldehyde and propionaldehyde. When this reaction was conducted using benzaldehyde 19,

meso-tetraphenylporphyrin (TPP) 20 was isolated (Scheme 6).12 In this reaction, pyrrole undergoes

electrophilic substitution at the α positions with an aldehyde to form the macrocyclic ring.

However, this reaction has many disadvantages. For instance, the reaction time was quite long and

the setup required to do the experiment was inconvenient as it required that the reaction be carried

20

out under anaerobic conditions in a sealed tube. In addition, purification of the products was very

difficult and poor yields were commonly obtained.12

Scheme 6 Rothemund Synthesis of TPP

Improvements on this approach were not reported until the 1960s. It was in 1967, that Adler

and Longo reported an improved method for the synthesis of meso-substituted porphyrins 20.13

This was accomplished by heating benzaldehyde and pyrrole in the presence of refluxing propionic

acid for 30 minutes (Scheme 7).13 The porphyrin product precipitated from the reaction mixture in

moderate yields of approximately 20%, which was a substantial improvement over Rothemund’s

method.12 The Adler and Longo’s method also required much shorter reaction times and could be

done under aerobic conditions. However, this approach still had a few drawbacks. Firstly, many

aldehydes could not tolerate the use of acidic conditions at elevated temperatures. Secondly, in

cases where the porphyrin products did not precipitate, difficulties were encountered in removing

side products that were also formed in these reactions.2

21

Scheme 7 Adler-Longo synthesis of TPP

In 1987, Lindsey et al. introduced a new methodology that greatly improved these

syntheses.14 In Lindsey’s method, the porphyrin synthesis was done in two steps instead of one.

The first step involves reaction of equimolar quantities of pyrrole and aldehyde in the presence of

an acid catalyst such as boron trifluoride etherate or trifluoroacetic acid (TFA) in dichloromethane

at room temperature. This results in the formation of a hexahydroporphyrin or porphyrinogen 21

(Scheme 8). Subsequently, in the second step oxidation of 21 with an electron deficient quinone,

p-chloranil or 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ), yields meso-substituted

porphyrin 22 in approximately 40% yield (Scheme 8).14 Due to the mild conditions of this

procedure, the Lindsey method allowed for the use of aldehydes with sensitive functionality that

could otherwise not be used under the relatively harsh conditions involved in previous routes to

meso-substituted porphyrins.

22

NH

RCHO

N

NH N

HN

R

R

R

R+

NH

NH HN

HN

R

R

R

RH

O

21

22

Scheme 8 Synthesis of meso-substituted porphyrins by Lindsey’s method

The methods developed by Rothemund, Adler, Lindsey and others are only suitable for

preparing symmetrical porphyrins and fall into the category of “One Pot Syntheses”. Since 1960,

many alternative routes to asymmetrical porphyrins have been developed including the ‘2+2’

MacDonald condensation, ‘3+1’ condensations, and the cyclization of open-chain tetrapyrroles.11

The ‘2+2’ condensations rely on dipyrromethane or dipyrromethene intermediates such as

23 and 24 (Figure 7). These precursors both have two linked pyrrole units and are differentiated

by the type of carbon linkage. Dipyrromethanes 23 have a saturated bridge, while dipyrromethenes

24 have an unsaturated bridge and are usually isolated as halide salts. Fischer made extensive use

of ‘2+2’ condensations of dipyrromethenes (Scheme 9) to synthesize asymmetrical porphyrins.10

23

Figure 7 Structures of Dipyrrolic Intermediates

For example, dipyrromethenes 25 and 26 condensed together in a succinic acid melt to

form deuteroporphyrin IX 27. Deuteroporphyrin 27 was a key intermediate in the total synthesis

of heme.11

Scheme 9 Fischer Porphyrin Synthesis

The use of dipyrromethanes was initially overlooked, in part because they were considered

to be too unstable to survive condensation reactions. However, in the 1960s, MacDonald was able

to demonstrate that dipyrromethanes are reasonably stable and excellent porphyrin precursors. In

24

MacDonald's “2+2” condensation method, a 5,5’-dipyrrylmethane dialdehyde 28 was condensed

with an α-unsubstituted dipyrrylmethane dicarboxylic acid 29 in the presence of an acid catalyst

to generate a porphodimethene intermediate 30, and subsequent air oxidation afforded the final

porphyrin product 31 (Scheme 10).15 Since the original report, the MacDonald reaction has been

one of the most commonly used methods for preparing meso-unsubstituted porphyrins. In addition,

a variation on MacDonald’s condensation was a crucial step in the total synthesis of chlorophyll

a.15

Scheme 10 MacDonald's “2+2” condensation reaction

Porphyrins can also be synthesized from oxophlorins 32, which are cross-conjugated

nonaromatic compounds that are the keto tautomers of meso-hydroxyporphyrins. Oxophlorin 32

25

can be synthesized using MacDonald ‘2+2’ conditions by reacting dipyrroketone dialdehydes 33

with dipyrromethanes 34 (Scheme 11).11 Although oxophlorins favor the keto tautomer over the

aromatic hydroxy tautomer, reaction of oxophlorins with acetic anhydride and pyridine traps the

hydroxy tautomer as the corresponding acetoxyporphyrin 35. Hydrogenation of 35 over palladium

yields porphyrinogen 36, and subsequent oxidation with DDQ affords the meso-unsubstituted

porphyrin 37 in reasonable yields (Scheme 11).11

Scheme 11 Synthesis of Porphyrin from an Oxophlorin

In the synthesis of porphyrins using previously mentioned methods, the introduction and

control of symmetry remains challenging. For instance, in the ‘2+2’ approach one of the precursors

must be symmetrical or mixture of isomers will be produced. In some cases, these isomeric

mixtures can be separated by careful column chromatography but the yields of individual

porphyrins will be low. An alternative ‘3+1’ version of the MacDonald condensation has been

26

developed that enables the synthesis of structures that cannot be prepared by the “2+2” approach,

although one of the intermediates must again be symmetrical. Johnson first reported a MacDonald

type “3+1” method for porphyrin synthesis in 1971.16 In this condensation reaction (Scheme 12),

tripyrrane dicarboxylic acids were reacted with dialdehydes to afford phlorins, which on

subsequent oxidation using DDQ or chloranil generated porphyrinoid structures. Johnson initially

used this approach to prepare oxa- and thiaporphyrins, but this methodology was later extended to

the synthesis of porphyrins and porphyrin analogues. This approach makes other types of

substitution patterns accessible, thus considerably expanding the library of porphyrin structures.16

In the synthesis of heteroporphyrins 38, a furan or thiophene dialdehyde 39 was condensed with a

tripyrrane 40 under acidic conditions and following oxidation yielded monoheteroporphyrins

(Scheme 12).16

Scheme 12 Synthesis of Oxaporphyrins and Thiaporphyrins

The key precursors in the ‘3+1’ MacDonald condensation are tripyrrolic intermediates

commonly known as tripyrranes. The first example of a tripyrrane 41 was synthesized by Kenner

and coworkers using 2-acetoxymethylpyrrole 42 and dipyrromethane 43.17 (Scheme 13). Sessler

et al. later reported a more convenient method for preparing tripyrranes in which an α-unsubstituted

27

pyrrole 44 reacted with two equivalents of an acetoxymethylpyrrole 45 in the presence of an acid

catalyst to generate 46. Hydrogenation of 46 yields tripyrrane dicarboxylic acids 47 (Scheme 14).18

Scheme 13 Synthesis of the First Tripyrrane

28

Scheme 14 Synthesis of Tripyrranes Reported by Sessler and Coworkers

Tripyrranes, in general, consist of three pyrrole rings linked by two methylene units.

However, analogous structures have been prepared where the central pyrrole has been replaced

with a heterocyclic or carbocyclic ring (Figure 8). Analogs of tripyrranes with benzene19 or

azulene20 units in place of the central pyrrole ring have been prepared and have been utilized in

the synthesis of highly modified porphyrinoid systems.19,21,22 The ‘3+1’ condensation has been

used to prepare a variety of porphyrin analogs including carbaporphyrins. This method is very

powerful and many different dialdehydes can be used to introduce exotic rings into the

porphyrinoid framework.23

29

Figure 8 Structures of Tripyrrane Analogues Used in ‘3+1’ Condensations

Porphyrins can also be synthesized by making use of tetrapyrrolic systems, such as bilanes,

bilenes and biladienes. However, tetrapyrrole cyclizations are heavily conditions dependent. The

tetrapyrrolic precursors are very reactive towards acid and can easily fragment or rearrange, hence

limiting the scope of these reactions. However, when appropriate conditions are used, tetrapyrroles

can remain intact and cyclize to generate isomerically pure porphyrins. For example,

dipyrromethanes 48 and 49 condense under mild acidic conditions to generate b-bilene 50.

Following cleavage of the tert-butyl esters, the b-bilene cyclizes in the presence of a mixture of

trifluoroacetic acid and trimethyl orthoformate, and following air oxidation generates isomerically

pure porphyrin 51 in good yields (Scheme 15).11

30

Scheme 15 Porphyrin Synthesis through a Tetrapyrrolic Intermediate

Reactivity of Porphyrins

Porphyrins are very reactive towards the formation of stable metal complexes. In fact, most

of the porphyrins found in nature are present as metalated derivatives. Porphyrins have been found

to form metal complexes with almost every metal and metalloid present in the periodic table;

however, the first row transition metals, such as iron, cobalt etc. tend to form metalloporphyrins

more readily compared to others.2 Porphyrin analogs such as carbaporphyrins also coordinate with

transition metal cations and in some cases stabilize unusual oxidation states. For instance,

31

benzocarbaporphyrins were shown to react with silver(I) acetate to form stable silver(III)

complexes, and related gold(III) derivatives (Figure 9) have also been prepared from gold(III)

acetate.24 Other carbaporphyrin-like macrocycles have also been shown to give similar

organometallic derivatives, e.g. tropiporphyrin (Figure 9).37

N

N

N

MN

N

N

Me

Et

Et

EtEt

Me

M

Ar

Ar

Ar

Ar

M = Ag(III) or Au(III) M = Ag(III)

Benzocarbaporphyrin derivatives Tropiporphyrin

Figure 9 Carbaporphyrins Stabilizing Metals in Rare Oxidation States

Porphyrins and metalloporphyrins can readily undergo electrophilic substitution reactions.

Metalloporphyrins are commonly substituted on their meso-carbons. However, if the coordinating

metals used are in electrophilic oxidation states (e.g. SnIV), or the free-base porphyrin is used, the

meso-carbons are deactivated towards electrophilic attack. This leads to electrophilic substitution

being favored at the β-positions of the pyrrolic subunits. These differences are illustrated in

Scheme 16 for the bromination of porphin.2

32

Scheme 16 Bromination of Porphyrins

Macrocycles Related to Porphyrins

Through the application of previously mentioned synthetic methods, such as ‘one pot

syntheses’ and ‘3+1’ condensations, many porphyrin-type macrocycles have been synthesized.

These porphyrinoids include contracted porphyrins such as triphyrin 5225 and corrole,2 and

expanded porphyrins such as sapphyrin 5326 and texaphyrin27 (Figure 10).

33

Figure 10 Structures of Contracted and Expanded Porphyrins

Phthalocyanines 54 are also structurally related to the porphyrins. These compounds are

important synthetic pigments that are typically blue or green in color. They were first investigated

by Sir Patrick Linstead, and like many other porphyrin-type structures, were found to be aromatic.

In phthalocyanine 54, each pyrrole ring is fused to a benzene ring and the pyrrole units are no

longer connected by methine bridges but are instead linked by nitrogen atoms (Figure 11). Analogs

of phthalocyanines have also been successfully synthesized. For instance, substitution of opposite

isoindole units with benzene rings gave dicarbahemiporphyrazine 55, although this structure no

longer retains aromatic properties.28

Figure 11 Structures of Phthalocyanine and Dicarbahemiporphyrazine

34

Applications

Synthetic porphyrins have a variety of applications, including as ligands in metal catalyzed

reactions and as photosensitizers in photodynamic therapy (PDT). By changing the chromophores

in porphyrin or its analogs, the macrocycle can be tuned for use in specific applications. For

instance, in PDT (a treatment for cancer), the specific properties of the photosensitizer has great

significance. PDT is a multistage process. Photosensitizer administration to the patient is carried

out, either systemically or topically, in the absence of light. When a sufficient amount of

photosensitizer appears in the diseased tissue, it is activated by exposure to light. The light dose

supplies sufficient energy to stimulate the photosensitizer and transfer the energy to molecular

oxygen. This results in the generation of reactive oxygen species such as singlet oxygen. The

singlet oxygen thus formed is destructive, and damages the malignant tumor cells.2 Porphyrins can

be used as photosensitizers in PDT because they are good at absorbing light and transferring the

absorbed energy to oxygen. They are excellent candidates as photosensitizers for PDT, as they

tend to accumulate in malignant cells and not healthy cells.29 In PDT, porphyrins or analogues that

show strong absorptions in the 650-800 nm range are preferred as photosensitizers. N-confused

porphyrins, carbachlorins and oxidized carbaporphyrins, for example, tend to have strong

absorptions in the far red and thus can be used for this type of application.23

Metalloporphyrins have been widely used in asymmetric catalysis, including

epoxidation,30 cyclopropanation,31 and Suzuki-Miyaura cross coupling.32 Development of

metalloporphyrins as catalysts was inspired by biological studies, such as the study of the

cytochrome P450 family of monooxygenases. These enzymes contain hemes (iron porphyrins) that

act as cofactors, and have been found to exhibit a broad range of functions. One of the most

significant functions of this class of enzymes is the elimination of foreign molecules in the body

35

through biotransformation. This family of enzymes is responsible for partial metabolism of the

majority of medicines consumed by humans.33 Another example of metalloporphyrin catalysts are

manganese porphyrins absorbed on a gold surface that can react with molecular oxygen from air

to carry out epoxidation of cis-stilbene to cis-stilbenoxide.

Aside from catalyzing organic reactions or being used as medicinal agents, porphyrins in

the modern world can also be utilized as chemosensors for nuclear waste. Expanded porphyrin

analogues have been found to bind with actinides and other heavy elements better than traditional

porphyrins. These expanded macrocyclic systems can also be used for the detection of hazardous

metals such as uranium or plutonium that are often used as a source for nuclear energy.34

Phthalocyanines, which are closely related to porphyrin macrocycles, have also been

shown to have numerous applications. They are also very stable compounds. For example, one of

the first phthalocyanines characterized by Linstead was stable up to 500 °C and in the presence of

concentrated sulfuric acid. Thus, these macrocycles and their analogues can be used as dyes for

inks, paper, and paint and have applications as lubricants and semiconductors.2

Porphyrins and phthalocyanines have also been shown to exhibit phosphorescence and

fluorescence properties.28 The phosphorescence and fluorescence properties of porphyrins can be

quenched by molecular oxygen. Thus, another application of porphyrins and phthalocyanines is

monitoring of oxygen levels used in various arenas from ecological, industrial, and medicinal

standpoints. These macrocycles can therefore serve the purpose of optical sensors for molecular

oxygen.28 The concentration of molecular oxygen in the gas or liquid phase is proportional to the

decay of phosphorescence in the porphyrin and phthalocyanine.28 The most common porphyrin

optical sensors are palladium(II) and platinum(II) metalloporphyrins. In addition, ruthenium(III)

and iridium(III) metalloporphyrins have also exhibited high phosphorescence at room temperature

36

and could possibly be used as alternative oxygen sensors in the near future.2

Carbaporphyrins

In order to gain a better understanding of the reactivity and characteristics of porphyrins,

chemists began to investigate core modified porphyrins. The study of core modified porphyrins

helps to reveal the effect that modification of a conjugated system has on its properties and its

characteristics. Systems with a carbon atom inserted into the core of the porphyrin, the so-called

carbaporphyrinoid systems, have been of particular interest. These systems have been widely

investigated and have shown unique properties, such as formation of organometallic derivatives

under mild conditions.35 Related systems include N-confused porphyrin 56,36

benzocarbaporphyrins 57,35 tropiporphyrins 58,37 azuliporphyrins 5938 and benziporphyrin 6039

(Figure 12).35

37

Figure 12 Monocarbaporphyrin Analogues

These porphyrin analogues can be subdivided into “true” carbaporphyrins and “modified”

carbaporphyrins. In true carbaporphyrins, at least one of the pyrrolic units is replaced with a

cyclopentadiene or indene unit. In modified carbaporphyrins, a carbocyclic ring system other than

a cyclopentadiene unit is used to replace a pyrrole moiety.35 For instance, by replacing a pyrrolic

unit within the porphyrin framework with azulene, cycloheptatriene, or benzene, new families of

porphyrin-like macrocycles were produced. True carbaporphyrins such as benzocarbaporphyrin

57 are fully aromatic structures, while benziporphyrins 60 do not have aromatic character, and

azuliporphyrins 59 fall midway between the two extremes.35

38

In benzocarbaporphyrins 57, porphyrin-like UV−vis spectra are retained and these

compounds are highly diatropic, as judged by proton NMR spectroscopy. However,

azuliporphyrins 59 are cross-conjugated and have considerably reduced diatropic character, while

tropiporphyrins 58 have intermediary properties due to the nonplanar nature of the seven-

membered ring.35 In principle, if more than one nitrogen atom in the cavity is replaced with carbon

atoms, this results in the formation of dicarbaporphyrins, tricarbaporphyrins and

tetracarbaporphyrins (Figure 13).41

Figure 13 Mono-, Di-, Tri- and Tetracarbaporphyrins

Carbaporphyrins can easily be synthesized by the “3+1” condensation approach.16 One of

the best studied classes of carbaporphyrins are the benzocarbaporphyrins. In benzocarbaporphyrins

57, one of the pyrrole units is replaced with an indene unit. These compounds are prepared by the

acid catalyzed condensation of indene dialdehyde 61 with tripyrrane 62 (Scheme 17)40. Following

39

oxidation with 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ), carbaporphyrin 57 was

isolated in > 40% yield.

Scheme 17 Synthesis of Benzocarbaporphyrin

Benzocarbaporphyrins can also be prepared from azuliporphyrin 63. In this method,

azuliporphyrins 63 undergo an oxidative ring contraction when exposed to tert-butyl

hydroperoxide under basic conditions to generate the fused benzene ring. This approach provides

an indirect route to tetraaryl benzocarbaporphyrins 64 (Scheme 18).42,43

40

Scheme 18 Synthesis of Tetraaryl Benzocarbaporphyrins

from Tetraarylazuliporphyrins

Tetraarylazuliporphyrin 63 can be easily prepared by reacting pyrrole, azulene and an

aromatic aldehyde together in the presence of boron trifluoride etherate. This route to

benzocarbaporphyrins is less effective than the ‘3+1’ methodology as two or more porphyrinoid

products result from the oxidative ring contractions. Nevertheless, this approach is currently the

only method available for synthesizing tetraaryl carbaporphyrins.42,43

41

Protonation of benzocarbaporphyrin 65 can lead to the formation of both mono and

dicationic species.44,45 Upon addition of acid, protonation initially occurs on the pyrrolenine

nitrogen to give the monocation 65H+. At higher acid concentrations, further protonation takes

place on the internal carbon to give dication 65H22+ (Scheme 19).45 The second protonation goes

to completion in 50% TFA-chloroform solution, and results in the generation of a new

chromophore. Proton NMR spectroscopy in 50% TFA-CDCl3 strongly suggests that the dicationic

species retains its aromatic characteristics. In azuliporphyrins, protonation of the internal carbon

atom has not been observed.35

Scheme 19 Protonation of Benzocarbaporphyrin

Benzocarbaporphyrins 66 were found to react with silver(I) acetate to give stable silver(III)

organometallic complexes 67 (Scheme 20).46,47 The silver complex retains porphyrin-like UV-vis

spectra, showing strong a Soret band at 437 nm, along with characteristic Q bands at 482, 518, 555

and 593 nm. The meso-protons show up at 9.89 and 10.06 ppm, thereby confirming that the silver

complex has strong diatropic character. Unlike the carbaporphyrin free base, the silver complex of

benzocarbaporphyrin 67 was near planar and indene unit was only tilted by 5.08°.46

42

NH

N

HN

Me

Et

R

REt

Me

66

N

N

N

Me

Et

R

REt

Me

67

Ag

Ag(OAc)

Scheme 20 Synthesis of Silver(III) Benzocarbaporphyrin

Reaction of meso-unsubstituted benzocarbaporphyrins 66 with gold(III) acetate gave low

yields of the related gold(III) complexes 68 (Scheme 21).47 However, much better yields were

obtained in the metalation of meso-tetraaryl benzocarbaporphyrins 64.35 Gold complexes of

benzocarbaporphyrins also have strongly aromatic characteristics and gave similar UV-visible

spectra to silver complexes 67.35

Scheme 21 Synthesis of Gold(III) Benzocarbaporphyrin

43

Recently, rhodium and iridium complexes of benzocarbaporphyrin 57 have also been

obtained in good yields (Scheme 22). Reaction of 57 with 1 equiv of di-μ-

chlorotetracarbonyldirhodium(I) in refluxing dichloromethane gave the rhodium(I) dicarbonyl

complex 69 in 90% yield. The proton NMR spectrum for 69 showed the presence of strong

diamagnetic ring currents. Upon refluxing 69 in pyridine, a hexacoordinate rhodium(III) complex

70 was generated in 55% yield. In addition, when 57 was reacted with [Ir(COD)Cl]2 in refluxing

pyridine, iridium(III) derivative 71 was generated in 22% yield. The X-ray structures for these

derivatives showed that the macrocycle was near planar and the bond lengths were consistent with

the aromatic structures.48

44

Scheme 22 Formation of Ir and Rh complexes of Benzocarbaporphyrin

Benzocarbaporphyrins also readily react with methyl iodide in the presence of potassium

carbonate in refluxing acetone to give alkylated products. In this reaction, it was observed that N-

methyl carbaporphyrin 72 was the major product along with some C-alkylated by-product 73

(Scheme 23). The major product of this reaction is a chiral system, whereas the minor C-methyl

product possesses a plane of symmetry. Although the presence of internal substituents may reduce

the planarity of these structures, these derivatives retain strongly diatropic characteristics.49

45

Scheme 23 Alkylation of Benzocarbaporphyrin

Unexpectedly, reaction of N-methyl carbaporphyrin 72 with palladium(II) acetate led to

the formation of a palladium(II) complex 74 where the methyl group has migrated from the

nitrogen atom to the carbon atom (Scheme 24).49 The NMR spectrum of 74 showed that the

complex is highly aromatic and the internal methyl resonance shifted upfield to -3.21 ppm.

Scheme 24 Unexpected Alkyl Group Migration

46

Reaction of benzocarbaporphyrin 57 with ferric chloride did not generate metal complexes

but instead gave various regioselective oxidation products.50, 51 When the reaction was carried out

using 500 equivalents of ferric chloride in a refluxing alcohol solvent, ketal derivatives 75 were

generated in high yields. These ketal derivatives were isolated in the monoprotonated form,

generally as the hydrochloride salt, and exhibited strong absorptions in the far red. Further

protonation of 75 using trifluoroacetic acid generated the related dicationic species. The proton

NMR spectra for 75 showed that these ketals retain strong diatropic characteristics and have a

plane of symmetry. For this system, the meso-protons in the proton NMR spectra showed up at

9.68 and 10.93 ppm. The far-red absorptions exhibited by these derivatives indicate that

carbaporphyrin ketals could have potential applications, for instance as photosensitizers in the

photodynamic therapy (PDT). In fact, 75.HCl has been shown to be an effective agent in the

treatment of leishmaniasis.

When the oxidation reaction was attempted using an aqueous solvent system for one hour,

a chloro derivative 76 was isolated in good yields. The chloro carbaporphyrin again retains

strongly aromatic characteristics, and the proton NMR spectrum of 76 showed that the meso-

protons were shifted to below 9.5 ppm. This is interesting as the X-ray crystal structure of the

chloro derivative shows that the indene subunit is significantly tilted away from the macrocyclic

plane due to the presence of a large internal Cl atom, hence substantially disturbing the planarity

of the molecule. In addition, the UV−vis spectrum of 76 shows a split Soret band at 421 and 434

nm, and Q bands at 521, 560, 608, and 665 nm, again signifying its porphyrin-like characteristics.51

When the reaction was carried out with ferric bromide instead of ferric chloride, under the

same reaction conditions, the corresponding bromo derivative of 76 was isolated in 7% yield. Even

though the larger bromine atom will undoubtedly further distort the macrocycle, the proton NMR

47

spectrum for 76b showed that the macrocycle surprisingly retained nearly all of its diatropic

characteristics. As reactions carried out in the presence of refluxing alcohols led to the formation

of ketals, it was anticipated that reactions using aqueous ferric chloride would afford the

corresponding dihydroxy products 77 or the related ketone. When the reactions with aqueous ferric

chloride were carried out for 16 h, a polar green byproduct 77 was isolated in 22% yield (Scheme

25).51 The proton NMR spectrum for this diketone showed that this species was nonaromatic. The

two carbonyl units gave rise to resonances at 173.6 and 197.6 ppm in the carbon-13 NMR

spectrum, but the IR spectrum showed the C=O stretching peaks at unusually low wave numbers.

The low frequencies can be attributed to the vinylogous amide nature of these bonds, which greatly

reduce the bond strengths.51

48

Scheme 25 Oxidation of Benzocarbaporphyrin with FeCl3

As carbaporphyrins exhibit new and interesting chemistry, steps toward the synthesis of

doubly modified systems were also undertaken. In the first successful synthesis of this type, indane

dialdehyde 78 was reacted with one molar equivalent of diethylpyrrole 44 in the presence of a

strong acid, and subsequent oxidation with FeCl3 generated the dibenzo opp-dicarbaporphyrin 79

in moderate yields (Scheme 26).52 Dibenzo opp-dicarbaporphyrin 79 proved to be somewhat

unstable in solution. Also, the low solubility of 79 in halogenated solvents made characterization

and in-depth reactivity studies very difficult. However, spectroscopic studies did show that the

49

system has strongly diatropic character as the proton NMR spectrum showed an upfield singlet for

the internal CH proton near -6 ppm, while the NH protons appeared near -5 ppm. The meso-

protons for this compound were strongly shifted downfield to around 9.8 ppm. The

dicarbaporphyrin also gave a UV-Vis spectrum that closely resembled true porphyrins with a

strong Soret band near 450 nm and a series of weak Q bands in the visible region.52

Scheme 26 Synthesis of Dibenzo opp-Dicarbaporphyrin

When 79 was protonated using a minimal amount of acid, it was observed that C-

protonation readily occurred to give monocation 79H+ (Scheme 27),52 and this caused a significant

change in the UV-Vis spectrum of 79. The Soret band was greatly diminished. In addition, a band

developed near 500 nm, and a strong absorption emerged near 750 nm. Proton NMR spectroscopy

showed that protonation had taken place at an internal carbon position, and this addition caused

little change in the diatropic nature of the system. The internal sp3 CH2 protons of the protonated

species were observed just above -4 ppm, while the CH proton was detected above -3 ppm. This

slight reduction in the upfield shift was balanced by a slight downfield shift of the meso protons to

50

+10 ppm. Diprotonation of dibenzo opp-dicarbaporphyrin was never observed even with TFA

concentrations in excess of 50%. This may be due in part to the increased steric hindrance that

would result from placing six hydrogens within the porphyrin core.52

Scheme 27 Protonation of Dibenzo opp-Dicarbaporphyrin

Carbachlorins

Chlorophylls are the central constituents in the engine of photosynthesis by which plants

harvest their energy. The structural nucleus of the chlorophylls is a dihydroporphyrin or chlorin

14 where one of the β,β-pyrrolic double bonds has been reduced.2 Studies of chlorin chemistry are

very important due to several reasons. These include the structure elucidation and total synthesis

of naturally occurring chlorophylls. In addition, chlorins show strong absorptions in the red region

of their UV-Vis spectra and synthetic chlorins have potential applications as photosensitizers in

PDT.

51

Even though one of the double bonds found in porphyrins has been reduced in chlorins, the

system still possesses an 18 π electron delocalization pathway. For this reason, chlorins are

aromatic compounds (Figure 14).53

Figure 14 Aromatic Delocalization Pathways in Porphyrin and Chlorin

Chlorins are isomers of nonaromatic phlorins and can potentially interconvert by

prototropic tautomerization (Scheme 28). In reactions that involve the intermediacy of meso-

hydrogenated tetrapyrroles such as porphyrinogens, tautomerization followed by 4e−/4H+

oxidation leads to the formation of chlorin. This type of tautomerization may be a source of chlorin

contamination in one-flask syntheses of porphyrins.53

52

Scheme 28 Phlorin-Chlorin Conversion

Although a great deal of work has been carried out in the field of synthetic chlorin

chemistry by various groups, few studies have explored the preparation of analogous systems such

as carbachlorins that may have equally useful properties. Carbachlorins are chlorin analogues in

which one of the core nitrogen atoms are replaced with a carbon atom (Figure 15).

53

Figure 15 Structures of Chlorin and Carbachlorin

In 1998, Lash and Hayes reported the synthesis of carbachlorins 80 using a “3 + 1” variant

on the MacDonald condensation (Scheme 29).54 In this synthesis, bicyclic dialdehydes 81 were

condensed with tripyrrane 62 in the presence of TFA and subsequent oxidation with DDQ led to

the formation of propanocarbachlorin 80a and propenocarbachlorin 80b in moderate yields. It was

interesting to note that these chlorin analogues retained aromatic characteristics, and in the proton

NMR spectra the inner CH protons showed up around -7 ppm. The UV-vis spectra of these

analogues were also very porphyrin-like, showing a strong Soret band at 404 nm and a series of

small absorptions that culminated in a medium sized Q band at around 650 nm. However, although

true chlorins are easily oxidized to give porphyrins, attempts to further oxidize chlorin analogues

80a or 80b to the corresponding carbaporphyrins 82 were unsuccessful.54

54

Scheme 29 Synthesis of Carbachlorins with Fused Five-membered Rings

In order to further explore the chemistry of carbachlorins, Lash and Hayes attempted to

synthesize porphyrinoid 83 by condensing 1,3-cyclopentanedicarbaldehyde 84 with a tripyrrane

62 (Scheme 30).55 However, no chlorinoid product was formed in these experiments. It was

speculated that the presence of β-pyrrolic substituents or the additional five-membered rings in

structure 82 aids in the generation of porphyrinoid products by altering the conformation of the

intermediates and facilitating the cyclization. For instance, indene dialdehyde 85 on treatment with

tripyrrane yields benzocarbaporphyrins 86 and cyclopentadiene trialdehyde 87 similarly reacted

with tripyrrane to give carbaporphyrin aldehyde 88, albeit in low yields (Scheme 30).55

55

Scheme 30 Synthesis of Carbaporphyrins and an Attempted Synthesis of a Carbachlorin

These results indicate that conjugated dialdehydes or aromatic dialdehydes such as 85 and

87 are more effective precursors in the “3 + 1” condensation reactions than aliphatic dialdehydes

56

such as 84. Therefore, in later studies, conjugated cyclopentene dialdehydes such as 89 were

targeted as precursors for carbachlorin formation (Scheme 31). Attempts to prepare 89 from a

related acetal were unsuccessful but a similar enol ether 90 could be obtained instead (Scheme 32).

This species was very unstable and was immediately used to prepare carbachlorin 83.

Scheme 31 Retrosynthesis of Desired Dialdehyde

In the reaction, 90 was directly condensed with tripyrrane dicarboxylic acid to generate

carbachlorin 83 in 11 – 16% yield (Scheme 32).55

57

Scheme 32 Synthesis of Carbachlorin

Carbachlorin 83 was aromatic in the free base form, displaying strong diatropic character

with the internal CH proton showing up at -6.93 ppm, while the meso-protons were downfield

appearing at 9.01 and 9.68 ppm. Since there were only two signals for the four meso-protons, the

carbachlorin system must possess a plane of symmetry. When carbachlorin 83 was treated with

trifluoroacetic acid, the corresponding monocation 83H+ was generated (Scheme 33).55 The proton

NMR spectrum for 83H+ indicated that the aromatic character was increased as the internal CH

now showed up at -7.00 ppm. Protonation also split the Soret band in the UV-visible spectrum of

carbachlorin 83 to give two different peaks showing up at 409 and 426 nm. Both the proton and

carbon-13 NMR spectra of 83H+ confirmed that the cation retained a plane of symmetry.55

58

Scheme 33 Protonation of Carbachlorin

Attempts to oxidize carbachlorins 80 (Scheme 29, page 54) to the corresponding

carbaporphyrins had been unsuccessful, but this may have been due in part to the presence of a

fused five-membered ring on carbachlorin 80. This unit would not favor the presence of the sp2

hybridized carbons generated upon oxidation due to the associated angle strain.15 In contrast, when

83 was heated with DDQ in toluene, oxidation readily occurred to give the corresponding

carbaporphyrin 91 (Scheme 34).55 Carbaporphyrin 91 is also highly aromatic and the proton NMR

spectrum gave a resonance for the internal CH at -6.91 ppm, while the NHs produce a broad peak

at -3.92 ppm. The meso-protons of 91 appeared downfield as two 2H singlets at 9.77 and 9.83

59

ppm. The proton and carbon-13 NMR spectra for 91 also confirmed that the macrocycle has a

plane of symmetry, and this result is consistent with tautomer 83 (Scheme 33).55

N

N

N

Et

EtMe

Et

H

Me

Et

H

N

N

N

Et

EtMe

Et

H

Me

Et

H HN

N

N

Et

EtMe

Et

H

Me

Et

H H

DDQ

Toluene

+ H+

- H+

+ H+- H+

91

91H+91H22+

N

N

N

Et

EtMe

Et

H

Me

Et

H

Scheme 34 Oxidation of Carbachlorin

The UV-vis spectrum of carbaporphyrin 91 was less porphyrin-like, showing two broad

absorptions in the Soret region at 377 and 421 nm, and broad Q bands at higher wavelengths. This

contrasts to previously mentioned benzocarbaporphyrins and formylcarbaporphyrins, which show

strong Soret bands and better-defined Q bands in their UV-vis spectra. The corresponding

monocation 91H+, generated by the addition of trace amounts of TFA, showed a strong Soret band

at 385 nm together with a broad absorption at 409 nm and Q bands at 563, 574 and 614 nm. Further

60

addition of TFA gave rise to a third species corresponding to the C-protonated dication 91H22+

(Scheme 34).55

It is interesting to note that the formation of benzocarbaporphyrin C-protonated dications

is only complete in approximately 50% TFA solutions, whereas diprotonation of 91 was essentially

complete in 5% TFA solutions. The proton NMR spectrum of 91H22+ showed the presence of a

very strong diatropic ring current in which the internal CH2 gave a resonance at -8.27 ppm while

the external cyclopentadiene protons showed up at 11.11 ppm and the meso-protons appeared

nearby as two 2H singlets at 11.00 and 11.45 ppm. These data clearly demonstrated that the

dication has enhanced aromaticity when compared to the free base form. However, the downfield

shift of the outer cyclopentadiene protons in 91H22+ is due to the relocation of the 18π electron

delocalization pathway through the five-membered carbon ring as shown in bold (Scheme 34).55

Carbaporphyrinoids such as 56, 57 and 58 readily react with silver(I) acetate to give

silver(III) organometallic derivatives.32,33 The formation of silver complexes of carbaporphyrins

57 occurs rapidly in mixtures of dichloromethane and methanol at room temperature in the

presence of 3 equiv. of silver(I) acetate. However, in the case of carbachlorin 83 the formation of

the silver(I) complex is rather slow under these conditions. With carbachlorin 83, the silver(III)

derivative 92 was isolated in 11% yield after almost 16 hr (Scheme 35). It was interesting to

observe that when the reaction was continued for several days, under same set of conditions, the

yield was raised to 31%. However, when 7 equivalents of silver(I) acetate was used, only low

yields (6%) of impure silver(III) carbaporphyrin 93 could be isolated.55

Further attempts to react carbaporphyrin 91 with silver(I) acetate led to decomposition

without forming any metalated product. These results clearly indicate that the exposed carbocyclic

ring is prone to oxidative degradation, and the data suggest that the presence of a fused benzene

61

ring in 57 is beneficial in stabilizing these structures and hence promotes metalation. The UV-vis

spectrum for 92 was porphyrin-like, showing a strong Soret band at 411 nm and a relatively strong

Q band at 599 nm. The proton NMR spectrum for 92 indicated that the complex retained strongly

aromatic properties as the meso-protons showed up downfield at 9.23 and 9.90 ppm and the

carbachlorin CH2CH2 unit gave a 4H singlet at 5.02 ppm.55 The presence of a plane of symmetry

was evident in the proton and carbon-13 NMR spectra, and the molecular formula was confirmed

by high resolution electron impact mass spectrometry.

Scheme 35 Formation of Silver(I) Complex of Carbachlorin

Benzocarbaporphyrin 57 as such does not give any isolatable products in reactions with

palladium(II) salts. However, as mentioned previously, the N-alkyl derivatives of

benzocarbaporphyrins were found to give palladium(II) complexes when reacted with

palladium(II) acetate in refluxing acetonitrile. This chemistry is interesting because the metalation

is associated with an alkyl group migration from the nitrogen to the internal carbon atom.35 A

similar alkylation-metalation sequence was conducted on carbachlorin 83.35

Reaction of 83 with methyl iodide and potassium carbonate in refluxing acetone for 16 h

gave the N-methyl derivative 94 in 34% yield (Scheme 36). Prolonged reaction times did not

62

improve the yield as unidentified side products were formed. The site of alkylation was deduced

from the proton and carbon-13 NMR spectra as the product shows a complete loss of symmetry.

For instance, the meso-protons for 94 showed up as four 1H singlets at 8.89, 9.04, 9.37 and 9.58

ppm.35

Scheme 36 Formation of N-methylated Carbachlorin

The alkylation does introduce steric crowding within the macrocyclic cavity but the NMR

data showed that the chlorin retained most of its diatropic character. In the proton NMR spectrum

of 94, the internal methyl group gave a 3H singlet at -4.27 ppm, while the inner CH produced a

resonance at -6.31 ppm. As the methyl group is too large to pass through the cavity, carbachlorin

94 is chiral and this leads to some complexity in the proton NMR spectrum. The chirality, makes

the six methylene units present in the structure all nonequivalent and the individual protons

diastereotopic, so in principle this could lead to twelve different resonances. However, due to a

significant amount of overlap in this structure, far fewer peaks resolve and the carbachlorin

CH2CH2 unit gave rise to only three multiplets appearing at 4.40 (1H), 4.54 (1H) and 4.79 ppm

(2H).55

The UV-vis spectrum for 94 in CH2Cl2 is somewhat broadened compared to the spectrum

for 83. The alkylated carbachlorin has a Soret band at 407 nm and less well-defined Q bands

63

showing up between 500 and 664 nm. These differences in the UV-visible spectrum presumably

result from distortion of the carbachlorin chromophore due to the presence of the internal methyl

group. Addition of TFA resulted in the formation of a cation 94H+ with a split Soret band at 415

and 428 nm, and again the spectrum was somewhat broadened compared to 94.55

Methylated chlorin 94 reacted with palladium(II) acetate in refluxing acetonitrile to give a

palladium complex. Column chromatography on silica gel gave two fractions with similar

polarities corresponding to metalated products from the N-methyl carbachlorin. It was anticipated

that the product of this chemistry would be N-methyl palladium(II) complex 95. Methyl group

migration would not be expected in this case because that would interrupt the macrocyclic

conjugation pathway. Unexpectedly, the main product was identified as palladium(II)

carbaporphyrin 96 where an oxidation had occurred in addition to the methyl group migration.

Palladium complex 96 was isolated in 17% yield (Scheme 37).55

64

`

Scheme 37 Unexpected Formation of Pd Complex

The remaining fraction appeared to consist primarily of the expected palladium(II) chlorin

95, but this could not be isolated in pure form. The samples were always contaminated with 96,

and it soon became apparent that 95 was slowly converting into 96. In order to speed up the

oxidation and improve the yield of 96, the crude reaction solution was diluted with

dichloromethane and stirred vigorously with an aqueous ferric chloride solution for 24 h.

Following work up, the yield of palladium(II) complex 96 was raised to 34%. It is likely that initial

oxidation affords the N-methyl palladium(II) carbaporphyrin 97 (Scheme 37)55 and this

subsequently undergoes a rearrangement to give 96. The mechanism of the rearrangement is not

65

known, but it could be a stepwise process involving the transient formation of a Pd-methyl bond.

Palladium complex 96 has regained a plane of symmetry and retains strongly diatropic

characteristics. The proton NMR spectrum showed the internal methyl group at -4.46 ppm, while

the meso-protons appeared downfield as two 2H singlets at 10.00 and 10.42 ppm. The

cyclopentadiene protons are directly connected to the 18π electron delocalization pathway and as

a result are deshelded to give a 2H singlet at 9.60 ppm. The UV-vis spectrum is less distinctive,

giving broad bands in the Soret region at 394 and 440 nm.55 The molecular formula for 96 was

confirmed by high resolution electron impact mass spectrometry.

Although some intriguing results have been obtained for carbachlorins, these studies have

been limited due to the difficulties encountered in preparing this system. Furthermore, an effective

methodology for synthesizing carbaporphyrins without fused aromatic rings has not been reported

previously. In this thesis, a new synthetic route for preparing carbaporphyrins and carbachlorins

has been developed. The new porphyrinoids, which are stabilized by electron-withdrawing ester

functionalities, were prepared in excellent yields from cyclopentanedialdehydes. The

carbaporphyrin system proved to be an excellent organometallic ligand, forming complexes with

nickel(II), palladium(II), silver(III) and gold(III).

66

CHAPTER II

SYNTHESIS OF CARBAPORPHYRINS

Introduction to Carbaporphyrin

The synthesis of macrocyclic structures related to porphyrins has attracted considerable

attention, in part because modified structures may enhance valuable properties that are exhibited

by regular porphyrins.12 Carbaporphyrinoid systems, where one or more of the core nitrogens have

been replaced by carbon atoms, have been widely investigated over the past 20 years. Although

detailed studies on benzocarbaporphyrins have been reported, carbaporphyrins without a fused

benzene unit have only been prepared in low yields, and an improved route to carbaporphyrins of

this type is needed. Specifically, porphyrinoid 98 with two ester functionalities attached to the

carbocyclic ring is an appealing target for synthesis. This represent a new type of

carbaporphyrinoid structure. The proposed methodology for preparing this system is modeled after

the synthesis of porphyrin analogues containing bicyclooctane units previously developed in Dr.

Lash’s lab.22

67

Results and Discussion

A retrosynthetic analysis suggested that carbaporphyrin 98 and a related chlorin could be

obtained from cyclopentane dialdehyde 99 and the known tripyrrane 62. It was speculated that

acid catalyzed “3+1” condensation of dialdehyde 99 with tripyrrane 62 in presence of

trifluoroacetic acid, following oxidation with DDQ, would afford the targeted carbaporphyrin

(Scheme 38).54 In order to utilize the “3+1” methodology, it was initially necessary to synthesize

tripyrrane 62. The preparation of this tripyrrane can be carried out in 12 synthetic steps, where two

key pyrrolic intermediates 100 and 44 needed to be generated (Scheme 39). The tripyrrane was

initially formed with benzyl ester protective groups but these can be removed by hydrogenolysis

over Pd/C to afford the required dicarboxylic acid 62 (Scheme 38).

Scheme 38 Retrosynthetic Analysis of Carbaporphyrin 98

68

Scheme 39 Retrosynthetic Analysis of the Tripyrrane Intermediate

In order to synthesize pyrrole 100, it was first necessary to prepare pyrrole ester 101.

Reaction of diethyl malonate with sodium nitrite and acetic acid afforded oxime 102 and this was

reduced under hydrogen over 10% Pd/C to give diethyl aminomalonate 103. Alkylation of 2,4-

pentanedione with ethyl iodide and potassium carbonate in refluxing acetone gave the ethyl

substituted diketone 104, and this was condensed with 103 in refluxing acetic acid to produce

pyrrole ethyl ester 101 (Scheme 40).

69

Scheme 40 Preparation of Pyrrole Ester 101 using a Knorr-Type Synthesis

Pyrrole ethyl ester 101 was transesterified using benzyl alcohol in the presence of sodium

benzyloxide to give the corresponding benzyl ester 105 (Scheme 41).

Scheme 41 Formation of a Pyrrole Benzyl Ester

70

Pyrrole 105 was then selectively oxidized with lead tetraacetate in acetic acid to afford the

required acetoxymethyl pyrrole precursor 100 in good yields (Scheme 42).

Scheme 42 Formation of an Acetoxymethylpyrrole

The second pyrrolic precursor 3,4-diethylpyrrole 44, was synthesized via the Barton-Zard

methodology. Initially, glycine ethyl ester hydrochloride and triethylamine were heated to reflux

in methyl formate to afford N-formylglycinate (Scheme 43). Subsequent dehydration with

phosphorus oxychloride and triethylamine at 0 ℃ then gave ethyl isocyanoacetate 106.

Scheme 43 Formation of Ethyl Isocyanoacetate 105

71

Besides ethyl isocyanoacetate, acetoxy nitrohexane 107 was also required as a precursor to

pyrrole 44. Initially, a related nitro alcohol was prepared by reacting propionaldehyde with

1-nitropropane in the presence of potassium fluoride (Scheme 44). Subsequent treatment with

acetic anhydride under acidic conditions then afforded the required acetoxy nitro hexane 107 in

96% yield. Barton-Zard condensation of acetoxy nitrohexane 107 with ethyl isocyanoacetate 106

in the presence of two equivalents of the base 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) resulted

in the formation of pyrrole ethyl ester 108. The crude pyrrole ester was then refluxed in a solution

of sodium hydroxide in ethylene glycol at 190 oC and this resulted in saponification and

decarboxylation to give diethyl pyrrole 44 in 48% yield (Scheme 44). However, the dialkyl pyrrole

was found to be moderately unstable at room temperature and was stored in the freezer for future

use.

72

Scheme 44 Preparation of 3,4-Diethylpyrrole via a Barton-Zard Synthesis

Treatment of two equivalent of 100 with 3,4-diethylpyrrole 44 and acetic acid in refluxing

ethanol gave tripyrrane dibenzyl ester in 92% yield. The benzyl ester protective groups were

cleaved by hydrogenolysis over 10% Pd/C to give the corresponding dicarboxylic acid 62 (Scheme

45).54 This product was relatively unstable and therefore had to be used within one week even

when it was stored in the freezer.

73

Scheme 45 Preparation of Tripyrrane 62

In addition to preparing the tripyrrane, it was necessary to develop a synthetic route to

cyclopentane dialdehydes such as 99. In an initial attempt, dimethyl acetylenedicarboxylate

(DMAD) 109 was reacted with cyclopentadiene 110 at room temperature to give the Diels-Alder

adduct 111 (Scheme 46).57

Scheme 46 Preparation of Bicyclic Adduct 111

74

Göksu and coworkers reported that seven bicyclo[2.2.1]heptenes, including 111, were

cleaved to cyclic dialdehydes in good yields by oxidation with KMnO4-CuSO4.5H2O.58 However,

all attempts to reproduce this procedure with norbornadiene derivative 111 failed to generate any

dialdehyde. Hence, an alternative synthetic method or a different precursor was needed. Donohoe

and coworkers reported an alternative two steps route to dialdehydes.59 In this synthetic route, a

solution of norbornadiene 111 in tert-butyl alcohol was oxidized to give 112 using KMnO4 and

NaOH in tert-butyl alcohol and water at 0 oC (Scheme 47).59 However, when we attempted to

prepare 112 using this methodology, the yields were low.

Scheme 47 Formation of Diol 112

Alternative synthetic routes for diol formation were investigated. However, all of the

conditions gave poor results. Some of the failed attempts to form diol 112 are shown in Scheme

48.58,60,61 In order to circumvent this problem, the use of an alternative norbornene intermediate

was considered.

75

Scheme 48 Unsuccessful Attempts to Synthesize Diol 112

Dimethyl fumarate (113), a readily available starting material, underwent a Diels-Alder

cycloaddition with cyclopentadiene to form the bicyclic adduct 114 (Scheme 49). Several different

procedures were attempted to carry out this transformation. Superior results were achieved using

a literature procedure by Beare and coworkers.62 In this procedure, freshly cracked

cyclopentadiene was gradually added to an aqueous suspension of dimethyl fumarate (Scheme

49).62 The mixture was stirred at room temperature for 150 minutes and then extracted with ethyl

acetate to give an excellent yield of 114. This norbornene derivative was then reacted with KMnO4

and NaOH in tert-butyl alcohol and water at 0 oC to give the related dialcohol 115 (Scheme 50).59

76

Scheme 49 Formation of Bicyclic Adduct 114

Scheme 50 Formation of Diol 115

The diol product 115 was isolated as a white crystalline solid in 73% yield. In the 1H NMR

spectrum of resulting product, two broad resonances for the OH protons could be identified (Figure

16).

77

0.51.01.52.02.53.03.54.04.55.05.56.06.57.07.5 ppm

3.703.753.803.85 ppm

Figure 16 500 MHz 1H NMR Spectrum of Fumarate Diol in CDCl3

Diol 115 could be oxidatively ring opened using potassium periodate as an oxidizing agent

to give dialdehyde 99 in 55% yield (Scheme 51).54 The mechanism for ring opening of the bicyclic

adduct to give a five-membered ring dialdehyde involved the formation of a transient cyclic

periodate derivative. Elimination of an iodate ion affords the dialdehyde (Scheme 52). Dialdehyde

99 thus generated was very unstable and decomposed rapidly even at low temperatures. For this

reason, it was immediately utilized in the preparation of carbaporphyrin.

Scheme 51 Formation of Dialdehyde 99

78

Scheme 52 Mechanism for Ring Opening

Dialdehyde 99 was condensed with tripyrrane dicarboxylic acid 62 in the presence of

TFA/CH2Cl2. Initial condensation would be expected to produce a hexahydrocarbaporphyrin such

as 116. Following oxidation with DDQ in refluxing toluene, carbaporphyrin 98 was isolated as the

major product along with some carbachlorin 117 as the minor product (Scheme 53).54

79

Scheme 53 Preparation of Carbaporphyrin 98

During purification using column chromatography, three bands were collected. The first

band to elute off the column was bright red, and corresponded to a small amount of porphyrin

byproduct. The second major band corresponded to carbaporphyrin 98 and following

recrystallization using chloroform and methanol, this product was isolated as purple crystals in

40% yield. A third green band was collected that corresponded to carbachlorin 117, and following

80

recrystallization from chloroform and methanol the dihydroporphyrinoid was obtained as a green

powder in 9% yield.

As expected, carbaporphyrin 98 proved to be a fully aromatic compound. The proton NMR

spectrum for 98 (Figure 17) in CDCl3 showed the presence of a strong diamagnetic ring current.

The internal CH proton was found at -6.27 ppm, while the meso-protons were shifted downfield

to give two 2H singlets at 10.24 and 9.51 ppm. As there were only two signals for the four meso-

protons, the results indicate that the molecule is symmetrical.

-6-5-4-3-2-111 10 9 8 7 6 5 4 3 2 1 0 ppm

3.853.903.954.00 ppm

Figure 17 500 MHz proton NMR Spectrum of Carbaporphyrin 98 in CDCl3

For the carbachlorin byproduct 117, the proton NMR spectrum in CDCl3 again showed the

presence of a strong diamagnetic ring current. The internal CH proton was found further upfield

at -7.14 ppm, while the meso-protons were still deshielded to between 9.85 and 9.48 ppm (Figure

81

18). The symmetry was maintained in the carbachlorin byproduct as can be seen in the figure

below.

-7-6-5-4-3-2-111 10 9 8 7 6 5 4 3 2 1 0 ppm

-7.1 -7.2 -7.3 ppm

9.69.8 ppm

Figure 18 500 MHz proton NMR spectrum of the Carbachlorin byproduct in CDCl3

The UV-Vis spectrum for the free base carbaporphyrin had a strong absorption in the Soret

region at 436 nm, followed by a series of four weaker absorptions between 520 and 720 nm, which

is characteristic aromatic porphyrinoids (Figure 19). In addition, the structure of carbaporphyrin

98 was also confirmed using X-ray crystallography (Figure 20).

82

Figure 19 UV-Vis spectrum of Carbaporphyrin 98 in Dichloromethane

0

2

4

6

8

10

12

300 400 500 600 700 800

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

83

Figure 20 POV-Ray generated ORTEP III Drawing of 98

The UV-vis spectrum for the free base carbachlorin byproduct 117, has a strong absorption

in the Soret region at 398 nm, followed by weaker absorptions at 490 nm and 588 nm. In addition,

a strong red shifted peak was displayed at 648 nm (Figure 21), a feature that is also seen for

tetrapyrrolic chlorins.

84

Figure 21 UV-Vis spectrum of Carbachlorin 117 in Dichloromethane

In principle, the monocation of 98 can be obtained when a trace amount of TFA is added.

However, when one drop of TFA was added to an NMR tube containing a solution of 98 in CDCl3,

the resulting NMR spectrum appeared to correspond to mixture of mono 98H+ and dicationic

species 98H2+ (Figure 22).

0

5

10

15

20

25

300 400 500 600 700

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

85

Figure 22 500 MHz proton NMR Spectrum of a Mixture of 98H+ and 98H22+ in CDCl3

When only 4 μL of TFA was added to the solution, the proposed monocation 98H+ (Scheme

54) could be observed by proton NMR and UV-Vis spectroscopy (Figure 23). The proton NMR

spectrum of 98H+ showed that the addition of TFA to the carbaporphyrin had slightly increased

the aromatic character of carbaporphyrin, as the meso-protons were shifted slightly further

downfield to 10.40 and 9.74 ppm (Figure 24). The appearance of a broad peak at -3 ppm

corresponds to the presence of an additional NH due to protonation. In addition, there was a

significant change in the UV-Vis spectrum with the addition of the trace amounts of acid (Figure

23).

-7-6-5-4-3-2-113 12 11 10 9 8 7 6 5 4 3 2 1 0 ppm

86

Scheme 54 Protonation of Carbaporphyrin 98

Figure 23 UV-Vis spectrum of Carbaporphyrin 98H+ in CH2Cl2

0

1

2

3

4

5

6

7

8

9

300 400 500 600 700 800

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

87

-6-5-4-3-2-112 11 10 9 8 7 6 5 4 3 2 1 0 ppm

10.0 ppm

3.23.43.63.84.04.24.4 ppm

-2.0 ppm

Figure 24 500 MHz proton NMR Spectrum of 98H+ in CDCl3

Further addition of TFA led to the formation of dication 98H22+ (Scheme 55). The proton

NMR spectrum of 98H22+ (Figure 25) showed a 2H singlet at -7.67 ppm due to the internal CH2,

and two 2H singlets at 11.80 and 11.00 ppm corresponding to the meso-protons. This not only

confirmed that diprotonation had taken place but also indicated that the aromatic character of the

conjugated system was further improved. In addition, the proton NMR spectrum for dication

98H22+ confirmed that the compound retained a plane of symmetry.

88

Scheme 55 Further Protonation of Carbaporphyrin 98H+

-7-6-5-4-3-2-113 12 11 10 9 8 7 6 5 4 3 2 1 0 ppm

-7.5 ppm

Figure 25 500 MHz 1H NMR Spectrum of 98H22+ in CDCl3.

89

In order to further understand the reactivity of carbaporphyrin 98, a few more experiments

were performed to observe how the exchange of protons takes place with the addition of acid. It

was interesting to note that addition of d-TFA to an NMR solution of 98 in CDCl3 showed rapid

exchange of the 21-H resonance and slow exchange of the meso-protons. As the solutions consisted

of a mixture of protonated species, it was necessary to extract the sample before NMR analysis

could be conducted. Two drops of d-TFA were added to the NMR tube and after 10 min the

solution was neutralized by addition of NaOD in D2O. The mixture was diluted with water,

extracted with dichloromethane and the organic solutions dried over sodium sulfate. Following

removal of the solvent, the residue was analyzed by proton NMR spectroscopy. The results showed

that the internal CH had undergone exchange but no other resonances were affected (Figure 26).

When the solution was exposed to d-TFA for 24 h, and the solution neutralized and extracted as

before, significant reduction in the intensities of the meso-proton resonances were observed. The

peak corresponding to the 5,20-protons was reduced by approximately 15%, while the 10,15-

proton resonance was diminished by >40% (Figure 27) The results indicate that meso-protonated

species such as monocations 98aH+ and 98bH+are energetically accessible and in equilibrium with

98, 98H+ and 98H22+ (Scheme 56). Similar results have previously been noted for

benzocarbaporphyrins.

90

Scheme 56 Proposed Equilibria of Different Cationic Species for Carbaporphyrin 98

91

-6-5-4-3-2-110 9 8 7 6 5 4 3 2 1 0 ppm

0.1

7

12

.29

5.6

73

.96

4.2

2

5.8

9

2.0

6

2.0

0

Figure 26 500 MHz proton NMR Spectrum of 98 Showing Exchange of the 21-H after 10 min

Exposure to d-TFA

-6-5-4-3-2-110 9 8 7 6 5 4 3 2 1 0 ppm

0.2

7

12

.84

5.8

3

4.1

2

4.2

1

6.0

0

1.1

2

1.7

0

Figure 27 500 MHz proton NMR Spectrum of 98 Showing Partial Exchange of the meso-

Protons after 24 hr Exposure to d-TFA

92

When carbachlorin 117 was exposed to a trace amount of TFA, a deep purple solution of

the corresponding monocation 117H+ (Scheme 57) could be observed by proton NMR (Figure 28)

and UV-Vis spectroscopy (Figure 29). The proton NMR spectrum for 117H+ showed that the

addition of acid to the carbachlorin system had slightly increased its aromatic nature, and although

the internal CH was still observed around -7 ppm, the meso-protons were shifted further downfield

to 10.30 and 9.84 ppm (Figure 28). The appearance of an additional broad peak at -6.0 ppm

confirmed the presence of an additional NH due to protonation. Unlike the related carbaporphyrins,

further addition of TFA to 117H+ did not give a diprotonated species as protonation at the inner

carbon would result in a disruption of macrocyclic aromatic character.

Scheme 57 Protonation of Carbachlorin byproduct 117H+

93

-7-6-5-4-3-2-112 11 10 9 8 7 6 5 4 3 2 1 0 ppm

9.810.010.2 ppm

-7.05 -7.10 ppm

-5.6 ppm

3.94.04.1 ppm

Figure 28 500 MHz 1H NMR Spectrum of 117H+ in CDCl3

94

Figure 29 UV-vis spectrum of 117H+ in 5% TFA-Dichloromethane.

Metalation of carbaporphyrin 98 was also investigated. As several carbaporphyrinoid

systems have been shown to form silver(III) derivatives, the synthesis of a silver(III) complex of

carbaporphyrin 98 was attempted. A solution of carbaporphyrin 98 in CH2Cl2 was added to a

suspension of silver(I) acetate in methanol, and the resulting mixture was stirred overnight at room

temperature. After column chromatography, a red fraction was collected and recrystallized using

chloroform and methanol to yield the organometallic complex 118 as a dark red powder in 60%

yield (Scheme 58).63 Silver complex 118 retains aromatic characteristics as indicated by proton

NMR spectroscopy (Figure 30). The meso-protons were deshielded showing up as two 2H singlets

at 9.91 and 8.79 ppm. However, due to metal complexation, there were no upfield resonances

beyond the TMS reference as the core CH and NH protons of the carbaporphyrin are no longer

present.

0

5

10

15

20

300 400 500 600 700

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

95

Scheme 58 Synthesis of Silver(III) Carbaporphyrin 118

Figure 30 500 MHz 1H NMR Spectrum of 118 in CDCl3.

In addition to the silver complex, the formation of a related gold(III) complex using

gold(III) acetate was attempted. Similar reaction conditions to those used for silver complex

formation were implemented. However, this reaction did not yield the gold complex. Different

96

reaction conditions were tried using the same solvent system but no metalated product could be

isolated. However, when 98 was refluxed with gold(III) acetate in pyridine, the solution rapidly

changed color from brown to dark red once it started to reflux. After 5 minutes of refluxing, it was

determined by TLC analysis that not all of the starting free base porphyrin 98 had reacted with the

gold(III) acetate. After 30 minutes of reaction time it was observed that most of the conversion has

taken place, and the reaction was stopped. Purification using column chromatography gave the

gold(III) complex 119 in 11% yield (Scheme 59). The derivative also retained aromatic

characteristics and gave a quite similar UV-vis spectrum to the silver complex.

Scheme 59 Formation of Gold(III) Complex of Carbaporphyrin

Previously, a benzocarbaporphyrin had been shown to react with methyl or ethyl iodide

and potassium carbonate in refluxing acetone to afford N- and C- alkylation products.49 Therefore,

it was of interest to see whether the new carbaporphyrin system, with two electron withdrawing

ester groups on adjacent carbons, could also be alkylated under similar conditions. It was

interesting to observe that when carbaporphyrin 98 was stirred with potassium carbonate and

methyl iodide in refluxing acetone solution for 30 minutes, an alkylation product was obtained in

97

up to 55% yield after purification (Scheme 60).64 However, unlike the alkylation of

benzocarbaporphyrin 57, only one monoalkylation product 120 was observed for this reaction.

Scheme 60 Formation of C-methylated Carbaporphyrin 120

In the proton NMR spectrum of 120 (Figure 31), the presence of singlet at -4.68 ppm (3H)

indicated that a methyl group had been introduced within the macrocyclic cavity. The symmetry

of the system was still maintained as there were only two signals for the meso-protons showing up

at 9.90 and 9.28 ppm. The spectroscopic data were consistent with a C-methyl derivative rather

than the alternative N-alkylation product.

98

-7-6-5-4-3-2-110 9 8 7 6 5 4 3 2 1 0 ppm

Figure 31 500 MHz 1H NMR Spectrum of 120 in CDCl3

When carbaporphyrin 98 was stirred with potassium carbonate using an excess of

iodomethane in acetone and the reaction was refluxed overnight, further reaction took place.

Following column chromatography, a major dialkylation product was isolated with methyl groups

on both the inner carbon and an adjacent nitrogen atom (Scheme 61).49 After recrystallization from

chloroform-methanol, the dimethylated carbaporphyrin 121 was isolated as shiny purple crystals

in 69% yield.

99

Scheme 61 Formation of Dimethylated Carbaporphyrin 121

In the proton NMR spectrum for the dimethylated carbaporphyrin, the internal methyl

groups showed up as two 3H singlets at -5.27 and -3.91 ppm. The four meso-protons appeared as

four separate singlets indicating a loss of symmetry in the structure. This loss of symmetry could

be explained if one of the methyl groups is attached to a pyrrole nitrogen that is flanking the

carbocyclic ring. In addition, as the methyl groups are too large to fit within the cavity, they could

both be placed to one side of the structure. It is also likely that the two methyl groups are placed

on opposite sides of the macrocycle. The resulting structure is chiral and would exist as a pair of

enantiomers. Although the presence of internal methyl groups may distort the structure, 121 is still

a strongly aromatic compound. As noted above, the internal methyl resonances are highly shielded,

while the meso-protons are relatively deshielded and show up at 8.90, 9.05, 9.52 and 9.54 ppm

(Figure 32).

100

Figure 32 500 MHz 1H NMR Spectrum of 121 in CDCl3

Addition of TFA to a solution of C-methylcarbaporphyrin 120 led to the formation of

monocation 120H�. The proton NMR spectrum for this species indicated a slight increase in

aromatic character is developed on protonation, as the internal methyl resonance is shifted upfield

to -5.41 ppm while the meso-protons are shifted downfield to 9.49 and 10.03 ppm (Figure 33).

101

-6-5-4-3-2-112 11 10 9 8 7 6 5 4 3 2 1 0 ppm

9.69.810.0 ppm

Figure 33 500 MHz 1H NMR Spectrum of 120H+ in CDCl3

Further addition of TFA led to the formation of a diprotonated dication 120H22+ showing

a doublet at -6.55 ppm due to protonation at the internal methyl carbon (Figure 34). The meso-

protons lie close to the TFA peak, showing up 11.08 and 11.7 ppm, indicating increased

diamagnetic character. It is worth noting that both the mono and dications maintained the plane of

symmetry.

102

Figure 34 500 MHz 1H NMR Spectrum of 120H22+ in CDCl3

Addition of TFA to a solution of dimethylcarbaporphyrin 121 led to the formation of

monocation 121H�(Figure 35). However, further addition of acid did not generate the related

dication. The proton NMR spectrum of 121H+ showed four singlets for the meso-protons at 9.92,

9.86, 9.46 and 9.45 ppm. The internal methyl resonances appeared at -3.54 and -4.91 ppm. The

results indicate that the protonated species has a slightly stronger diatropic ring current than the

free base form 121.

-6-5-4-3-2-112 11 10 9 8 7 6 5 4 3 2 1 0 ppm

103

-5-4-3-2-111 10 9 8 7 6 5 4 3 2 1 0 ppm

9.59.69.79.89.9 ppm

Figure 35 500 MHz 1H NMR Spectrum of 121H+ in CDCl3

Metalation of N-methylated benzocarbaporphyrin with palladium(II) acetate in refluxing

acetonitrile has previously been shown to give a palladium(II) organometallic derivative.64 The

metalation reaction in benzocarbaporphyrins is associated with an alkyl group migration from the

nitrogen to the inner carbon atom. However, in 120 the methyl group is already attached to the

internal carbon and no rearrangement would be necessary to form an analogous palladium

complex. In fact, reaction of 120 with palladium(II) acetate is reflexing acetonitrile gave the

organometallic derivative 122 in 63% yield (Scheme 62).49 Although, the conjugation pathway in

the metal complex has been altered, the macrocycle still remains fully aromatic.

104

Scheme 62 Metalation of C-methylcarbaporphyrin with Palladium(II) Acetate

The proton NMR spectrum of 122 (Figure 36) gave two 2H singlets at 9.85 and 10.81 ppm

in the downfield region for the four meso-protons, confirming that compound 122 has a plane of

symmetry. Moreover, the presence of a singlet at -4.12 ppm (3H) indicated that the internal methyl

group was strongly shielded. As the symmetry of the alkylation product after complexation is still

maintained, the results indicate that no rearrangement has taken place. The UV-Vis spectrum for

the Pd complex shows a strong Soret band at 306 nm and Q bands at 410, 460, 602 and 639 nm

(Figure 37) that further support its aromatic character. The structure of palladium complex 122

was also confirmed using X-ray crystallography (Figure 38).

105

-8-7-6-5-4-3-2-114 13 12 11 10 9 8 7 6 5 4 3 2 1 0 ppm

Figure 36 500 MHz 1H NMR Spectrum of 122 in CDCl3

Figure 37 UV-vis Spectrum of Palladium(II) Carbaporphyrin in Chloroform

0

0.5

1

1.5

2

2.5

3

3.5

4

300 400 500 600 700 800 900

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

106

Figure 38 POV-Ray rendered ORTEP III Drawing of 122

Nickel complexes of carbaporphyrins have not been previously reported and attempts were

made to prepare a similar complex from nickel(II) acetate. Under the same set of conditions, no

formation of a nickel complex was observed. However, when 120 was refluxed with nickel(II)

acetate in DMF, the solution rapidly turned from brown to dark green as soon as it started to reflux.

After 5 minutes of refluxing, it was determined by TLC analysis that all of the starting

carbaporphyrin 120 had reacted with the nickel(II) acetate. After column chromatography, a green

colored fraction was obtained that corresponded to nickel(II) complex 123. After recrystallization

from methanol and chloroform, the pure complex was obtained in 71% yield. (Scheme 63).

107

Scheme 63 Metalation of C-methylcarbaporphyrin with Nickel(II) Acetate

The proton NMR spectrum of 123 (Figure 39) showed two singlets at 9.59 and 10.56 ppm

in the downfield region for the four meso-protons, suggesting that compound 123 has retained the

plane of symmetry. Moreover, the presence of a singlet at -3.89 ppm (3H) indicated that an internal

methyl group was still present. Clearly, the macrocycle still retained fully aromatic characteristics.

The structure was also confirmed by X-ray crystallography (Figure 40).

108

-10-9-8-7-6-5-4-3-2-114 13 12 11 10 9 8 7 6 5 4 3 2 1 0 ppm

NN

N

Et

Me

MeO2C

MeO2C Me

Et

Et

Et

Ni

Me

Figure 39 500 MHz 1H NMR Spectrum of 123 in CDCl3

Figure 40 POV-Ray rendered ORTEP III Drawing of 123

109

Conclusions

A cyclopentane dialdehyde was prepared in two steps from the Diels–Alder adduct of

dimethyl fumarate and cyclopentadiene. The ‘3+1’ condensation of dialdehyde 99 and tripyrrane

dicarboxylic acid 62 led to the formation of carbaporphyrin 98 in 40% yield together with some

carbachlorin 117 byproduct. This newly synthesized carbaporphyrin system proved to be highly

aromatic. Protonation of this system can readily occur, forming both mono- and dications.

Alkylation and metalation products were also synthesized and characterized.

Carbachlorin 117 can also readily form a monocation 117H+ when TFA is added, but no

dication formation was observed. Reaction between carbaporphyrin 98 and silver(I) acetate gave

a silver(III) organometallic complex. Similarly, reaction between carbaporphyrin 98 and gold(III)

acetate afforded a gold(III) organometallic complex, albeit in low yield. Alkylation of

carbaporphyrin 98 gave a C-methylated product 120 when refluxed with methyl iodide for 1 hr,

but a dimethylated product predominated when the reaction was carried out overnight. Protonation

of the monoalkylated carbaporphyrin with TFA gave both monocationic and dicationic species,

but the dimethylated species only formed a monocation. Metalation of 120 with palladium(II)

acetate afforded palladium(II) carbaporphyrin, while reaction with nickel(II) acetate afforded a

related nickel(II) carbaporphyrin in good yields.

The straightforward synthesis of carbaporphyrin 98 will allow further studies on this

system to be conducted. It is anticipated that the ester moieties can be reduced to give

carbaporphyrin alcohols. Initial investigations with LiAlH4 led to decomposition, but reaction with

diisobutylaluminium hydride (DIBAL-H) selectively reduced one of the ester units. At -70 ℃, the

related monoaldehyde 124 was generated but at room temperature monoalcohol 125 was formed

(Scheme 64). So far attempts to reduce both ester moieties have been unsuccessful.

110

Scheme 64 Reduction of Carbaporphyrin 98

Reduction to diol 126 could enable novel cycloaddition chemistry to be investigated.

Conversion to the related dibromide 127 might be accomplished by reaction with phosphorus

tribromide. Treatment with sodium iodide may result in the formation of a transient diene 128 that

could undergo a Diels–Alder cycloaddition with suitable dienophiles such as dimethyl

acetylenedicarboxylate to generate 129 (Scheme 65). This chemistry will make new classes of

carbaporphyrin derivatives accessible.

111

Scheme 65 New Class of Carbaporphyrin Derivatives

Experimental

All chemicals were purchased from Sigma Aldrich or Acros Organic. Grade III neutral

alumina and silica gel were used to perform column chromatography. The 1H and 13C-NMR

spectra were obtained either using Bruker Avance III 500 MHz NMR spectrometer or Bukner 400

MHz NMR spectrometer at 302 K. Chemical shifts were recorded, relative to CDCl3 (residual

chloroform at δ 7.26 ppm) in proton NMR spectra and the CDCl3 triplet at δ 77.23 ppm in 13C-

NMR spectra, in parts per million (ppm). 2D experiments were performed by using standard

112

software. The UV-vis spectra for the compounds were collected by Cary 100 Bio

spectrophotometer. Melting point were gathered using a Mel-Temp apparatus. High-resolution

mass spectra (HRMS) were carried out by using a double focusing magnetic sector instrument

Mass spectrometry data was obtained from the Mass Spectral Laboratory, School of Chemical

Sciences, University of Illinois, Urbana-Champaign.

Dimethyl bicyclo[2.2.1]hept-5-ene-2-endo-3-exo-dicarboxylate (114)

Freshly distilled cyclopentadiene (6.6 ml, 78 mmol) was added to a suspension of dimethyl

fumarate (10 g, 69.3 mmol) in water (400 mL) and the suspension was stirred vigorously for 120

minutes. Ethyl acetate (150 mL) was added and the phases were separated, the organic phase was

dried over Na2SO4 and the solvent evaporated to give the title compound (10.7 g, 50.9 mmol, 73%)

as a colorless oil. 1H NMR (400 MHz, CDCl3): δ 1.46 (1H, dq, J = 8.8, 1.7 Hz), 1.62 (1H, br t, J

= 8.8 Hz), 2.69 (1H, dd, J = 4.5, 1.7 Hz), 3.11-3.14 (1H, m), 3.25-3.28 (1H, m), 3.38 (1H, t, J =

4.1 Hz), 3.65 (3H, s), 3.72 (3H, s), 6.07 (1H, dd, J = 5.6, 2.8 Hz), 6.28 (1H, dd, J = 5.6, 3.1 Hz).

13C NMR (100 MHz, CDCl3): δ 45.9, 47.3, 47.5, 47.8, 48.1, 52.0, 52.3, 135.4, 137.8, 173.9, 175.1

113

Dimethyl 5,6-dihydroxybicyclo [2.2.1] heptane-2,3-dicarboxylate (115)

To a solution of norbornene 114 (10.83 g, 51.5 mmol) in tert-butyl alcohol (200 mL) and water

(50 mL) at 0 °C was added dropwise a solution of potassium permanganate (12.0 g, 76.0 mmol)

and sodium hydroxide (2.61 g, 65.3 mmol) in water (250 mL). After the addition was complete,

the reaction mixture was stirred for 20 minutes and then quenched using a saturated aqueous

solution of sodium metabisulphite (Na2S2O5) until the solution turned colorless. The mixture was

filtered and the tert-butyl alcohol was removed from the filtrate in vacuo. The solution was

extracted with ethyl acetate (4 × 100 mL) and the combined organic layers dried over MgSO4,

filtered and concentrated in vacuo to furnish the diol 115 (6.85 g, 28.0 mmol, 55%) as a white

crystalline solid, mp 84-86 oC (lit. mp59 81-84 oC). 1H NMR (400 MHz, CDCl3): δ 1.44 (1H, dp, J

= 11.0, 1.5 Hz), 1.89 (1H, dq, J = 11.0, 1.5 Hz), 2.50 (1H, br s), 2.55-2.57 (1H, m), 2.62 (1H, br),

2.73 (1H, br s overlapping with dd, J = 5.7, 1.3 Hz), 3.20 (1H, t, J = 5.1 Hz), 3.70 (3H, s), 3.71

(3H, s), 3.76-3.79 (1H, br m), 3.86-3.89 (br m). 13C NMR (100 MHz, CDCl3): δ 31.9, 45.0, 46.4,

46.7, 48.5, 52.4, 52.5, 70.6, 73.7, 173.1, 174.2. HRMS (EI): calcd for C11H16O6 + H 245.1025,

found 245.1027.

114

Dimethyl 3,5-diformylcyclopentane-1,2-dicarboxylate (99)

The foregoing diol 115 (4.15 g, 16.9 mmol) was mixed with water (5.0 ml), ether (10 ml) and

potassium periodate (4.66 g, 20.3 mmol), and the reaction was stirred at room temperature for 1

hr. The excess potassium periodate was filtered off and the two phases were separated. The water

layer was back extracted twice with ether. The ether layers were combined, dried over Na2SO4,

and the ether removed under reduced pressure to gave dialdehyde as a clear oil 99 (2.08 g, 8.59

mmol, 50%). The unstable dialdehyde was not further purified but used immediately to prepare

carbaporphyrin. 1H NMR (400 MHz, CDCl3): 2.17-2.26 (1H, m), 2.46 (1H, dt, J = 13.8, 5.9 Hz),

3.05-3.10 (1H, m), 3.26-3.32 (1H, m), 3.56-3.65 (2H, m), 3.70 (3H, s), 3.76 (3H, s), 9.65 (1H, d,

J = 0.8 Hz), 9.68 (1H, d, J = 0.9 Hz). HRMS (EI): calcd for C11H14O6 + H 243.0869, found

243.0871.

115

8,12,13,17-Tetraethyl-2,3-di(methoxycarbonyl)-7,18-dimethyl-21-carbaporphyrin (98)

Tripyrrane dicarboxylic acid 62 (100 mg, 0.22 mol) was stirred with TFA (1 ml) under an

atmosphere of nitrogen for 2 minutes in a 250 ml round bottom flask. Dichloromethane (99 ml)

was added, followed immediately by dialdehyde 99 (56 mg, 0.23 mol), and the mixture was stirred

under nitrogen overnight. The mixture was neutralized by the addition of triethylamine and the

solvent removed under reduced pressure. The residue was taken up in toluene (50 ml), DDQ (150

mg, 0.66 mmol) was added and the resulting mixture was stirred under reflux for 1 hr. The solvent

was removed under reduced pressure, then the residue was dissolved in CH2Cl2 and washed with

water. The solution was concentrated under reduced pressure and the residue chromatographed on

grade III neutral alumina column eluting with CH2Cl2 and hexane in a ratio of 60:40. The product

was collected as a dark brown fraction that was the second fraction eluting from the column.

Recrystallization from chloroform and methanol gave carbaporphyrin (52 mg, 0.091 mmol, 42%)

as purple crystals, mp 234-236oC; UV-Vis (1% Et3N-CH2Cl2): λmax (log ε) 368 (4.73), 436 (5.00),

526 (4.04), 566 (4.24), 648 (3.31), 717 nm (3.43). UV-Vis (5% DBU-CH2Cl2): λmax (log ε) 361

(4.59), 450 (4.80), 530 (sh, 3.55), 572 (sh, 3.80), 625 nm (4.01). UV-Vis (100 eq TFA-CH2Cl2):

λmax (log ε) 315 (4.46), 392 (4.77), 446 (4.91), 549 nm (4.05). UV-Vis (50% TFA-CH2Cl2): λmax

(log ε) 315 (4.15), 407 (5.17), 425 (5.12), 549 (3.79), 597 (3.96), 656 nm (3.27). 1H NMR (500

116

MHz, CDCl3): δ -6.27 (1H, s), -3.12 (2H, br), 1.77-1.83 (12H, 2 overlapping triplets), 3.51 (6H,

s), 3.84 (4H, q, J = 7.6 Hz), 3.95 (4H, q, J = 7.6 Hz), 4.27 (6H, s), 9.51 (2H, s), 10.24 (2H, s). 13C

NMR (125 MHz, CDCl3): δ 11.5, 17.3, 18.4, 19.5, 19.9, 52.6, 94.6, 100.7, 107.2, 130.5, 131.6,

135.5, 136.0, 137.7, 139.4, 145.4, 156.1, 167.6. 1H NMR (500 MHz, 4µl TFA-CDCl3): δ -5.95

(1H, s), -3.14 (1H, v br), -1.97 (2H, s), 1.65 (6H, t, J = 7.6 Hz), 1.80 (6H, t, J = 7.6 Hz), 3.44 (6H,

s), 3.90-4.00 (8H, 2 overlapping quartets), 4.26 (6H, s), 9.74 (2H, s), 10.40 (2H, s). 13C NMR (125

MHz, 4µl TFA-CDCl3): δ 11.7, 16.6, 17.4, 19.78, 19.81, 53.2, 94.2, 111.8, 113.8, 132.7, 134.1,

138.5, 140.2, 140.4, 140.5, 141.5, 142.9, 167.3. 1H NMR (500 MHz, 50% TFA-CDCl3): δ -7.65

(1H, s), 1.82 (12H, t, J = 7.6 Hz), 3.79 (6H, s), 4.22-4.32 (8H, m), 4.69 (6H), 11.10 (2H, s), 11.86

(2H, s). 13C NMR (125 MHz, 50% TFA-CDCl3): δ 11.7, 16.3, 17.4, 20.6, 29.2, 55.5, 105.1, 115.0,

142.7, 144.1, 144.8, 146.1, 147.1, 148.2, 148.2, 148.3, 149.0, 149.3, 167.3. HRMS (ESI): calcd

for C35H39N3O4 + H 566.3019, found 566.3047.

117

trans–8,12,13,17-Tetraethyl-2,3-di(methoxycarbonyl)-7,8-dimethyl-2,3-dihydro-21-

carbaporphyrin (117)

A third green fraction from the previous procedure was evaporated and the residue recrystallized

from chloroform-methanol to give the carbachlorin (12 mg; 0.02 mmol, 9%) as a green powder,

mp 232-234 oC; UV-Vis (5% DBU-CH2Cl2): λmax (log ε): 398 (4.86), 437 (4.59), 450 (sh, 4.52),

464 (sh, 4.44), 569 (sh, 3.74), 587 (sh, 3.75), 619 (sh, 3.90), 645 (sh, 3.98); UV-Vis (CH2Cl2): λmax

(log ε): 398 (5.31), 490 (4.17), 588 (3.41), 645 (4.23); UV-Vis (5% TFA-CH2Cl2): λmax (log ε):

348 (4.21), 405 (5.24), 422 (5.15), 527 (3.84), 544 (sh, 3.84), 582 (3.88), 635 (3.91); 1H NMR

(500MHz, CDCl3): δ -7.14 (1H, s, 21-H), 1.82 (6H, t, J = 7.7 Hz), 1.86 (6H, t, J = 7.7 Hz), 3.53

(6H, s), 3.92 (6H, s), 3.98 (4H, q, J = 7.7 Hz), 4.06 (4H, q, J = 7.7 Hz), 6.34 (2H, s), 9.48 (2H, s),

9.85 (2H, s). 13C NMR (125 MHz CDCl3): δ 11.5, 17.6, 18.8, 19.8, 20.1, 53.0, 56.0, 96.6, 97.9,

119.7, 130.0, 133.0, 135.7, 138.0, 139.6, 143.4, 150.5, 173.7. 1H NMR (500 MHz, TFA-CDCl3):

δ -6.96 (1H, s), -6.00 (1H, br s), -5.56 (2H, br s), 1.72 (6H, t, J = 7.7 Hz), 1.82 (6H, t, J = 7.7 Hz),

3.53 (6H, s), 4.01 (6H, s), 4.12 (4H, q, J = 7.7 Hz), 4.19 (4H, q, J = 7.7 Hz), 6.34 (2H, s), 9.84

(2H, s), 10.30 (2H, s). 13C NMR (125 MHz, TFA-CDCl3): δ 11.4, 16.6, 17.5, 20.2, 20.3, 54.8, 56.9,

94.9, 103.3, 128.4, 134.8, 135.2, 136.2, 141.7, 142.3, 142.8, 143.3, 176.1. HRMS (ESI): Calcd for

C35H41N3O4 + H 568.3175, found 568.3181.

118

(8,12,13,17-Tetraethyl-2,3-di(methoxycarbonyl)-7,18-dimethyl-21-carbaporphyrinato)silver(III)

(118)

A solution of carbaporphyrin 98 (10.0 mg, 0.017 mmol) in dichloromethane (10 mL) was added

to silver(I) acetate (11.0 mg, 0.066 mmol) dissolved in methanol (2.5 mL), and the mixture was

stirred at room temperature overnight. The mixture was washed with water, and the solvent was

removed under reduced pressure. The residue was chromatographed on grade III neutral alumina

eluting with dichloromethane and hexane in a ratio of 60:40. A deep red fraction was collected and

recrystallized from chloroform-methanol to give the silver(III) complex (7.2 mg, 0.011 mmol,

60%) as a red powder, mp >300oC; UV-Vis (CH2Cl2): λmax (log ε): 309 (4.06), 352 (4.29), 376

(4.29), 407 (4.44), 449 (4.64), 515 (3.63), 567 (4.13); 1H NMR (500 MHz, CDCl3): δ 1.59 (12H,

t, J = 7.7 Hz), 3.31 (6H, s), 3.59 (4H, q, J = 7.7 Hz), 3.64 ( 4H, q, J = 7.7 Hz) 4.40 (6H, s), 8.79

(2H, s), 9.91 (2H, s); 13C NMR (125 MHz, CDCl3): δ 11.5, 17.3, 18.1, 19.4, 20.0, 52.6, 96.1, 110.2,

118.3, 131.1, 134.51, 134.57, 137.2, 138.3, 139.6, 140.9, 167.5. HR MS (EI): calcd for

C35H36AgN3O4 669.1757, found 669.1752.

119

(8,12,13,17-Tetraethyl-2,3-di(methoxycarbonyl)-7,18-dimethyl-21-carbaporphyrinato)gold(III)

(119)

A solution of carbaporphyrin 98 (12 mg, 0.021 mmol) in pyridine (10 mL) was added to gold(III)

acetate (12 mg, 0.032 mmol) dissolved in pyridine (2.0 mL), and the mixture was refluxed for 30

minutes. The mixture was washed with water, and the solvent removed under reduced pressure.

The residue was chromatographed on Grade III neutral alumina eluting with dichloromethane and

hexane in 60:40. A red fraction was collected and recrystallized from chloroform-methanol to give

the gold(III) complex (2.6 mg, 0.004 mmol, 16%) as a red powder, mp 260-262 oC; UV-Vis

(CH2Cl2): λmax (log ε) 348 (4.38), 373 (4.35), 403 (4.41), 441 (4.43), 474 (sh, 3.83), 513 (3.76),

557 (4.41); 1H NMR (500 MHz, CDCl3): δ 1.63-1.68 (12H, t, J = 7.7 Hz), 3.36 (6H, s), 3.69 (4H,

q, J = 7.7 Hz), 3.74 (4H, q, J = 7.7 Hz), 4.36 (6H, s), 9.02 (2H, s), 9.99 (2H, s). 13C NMR (125

MHz, CDCl3): δ 11.5, 17.3, 18.0, 19.4, 19.8, 52.5, 96.3, 111.4, 122.1, 134.07, 134.14, 135.1, 136.2,

137.3, 138.2 140.9, 167.5. HR MS (ESI): calcd for C35H36AuN3O4 760.2450, found 760.2460.

120

8,12,13,17-Tetraethyl-2,3-di(methoxycarbonyl)-7,18,21-trimethyl-21-carbaporphyrin (120)

Carbaporphyrin 98 (12.0 mg, 0.021 mmol) was stirred with potassium carbonate (12 mg) and

methyl iodide (6 drops) in acetone (18 mL) under reflux for 1 h. The mixture was washed with

water and evaporated under reduced pressure. The residue was run through a grade III neutral

alumina column, eluting with dichloromethane and hexane in a ratio of 60:40, and two colored

fractions were collected. The later major red fraction was evaporated under reduced pressure and

the residue recrystallized from chloroform-methanol to yield the C-methylcarbaporphyrin (9.1 mg,

0.016 mmol, 73%) as purple powder, mp 298-300oC; UV-Vis (CH2Cl2): λmax (log ε) 312 (4.17),

428 (4.67), 537 (sh, 3.64), 580 (3.92), 647 nm (sh, 3.33). UV-Vis (1% Et3N-CH2Cl2): λmax (log ε)

408 (4.64), 431 (sh, 4.61), 468 (sh, 4.45), 576 (sh, 3.73), 649 (3.90), 701 nm (sh, 3.37). UV-Vis

(5% DBU-CH2Cl2): λmax (log ε) 405 (4.67), 432 (sh, 4.62), 467 (sh, 4.50), 564 (sh, 3.73), 648

(3.97), 700 nm (sh, 3.38). UV-Vis (5 equiv TFA-CH2Cl2): λmax (log ε) 316 (4.16), 411 (4.76), 556

(sh, 3.65), 581 (3.75), 629 nm (sh, 3.45). UV-Vis (25% TFA-CH2Cl2): λmax (log ε) 320 (3.78), 414

(5.08), 430 (sh, 4.89), 550 (3.46), 606 nm (3.79). UV-Vis (50% TFA-CH2Cl2): λmax (log ε) 316

(3.97), 413 (5.11), 432 (4.92), 529 (sh, 3.50), 549 (sh, 3.48), 598 (3.80), 620 nm (3.80). 1H NMR

(500 MHz, CDCl3): δ -4.68 (3H, s), -3.19 (2H, v br), 1.76-1.80 (12H, 2 overlapping triplets), 3.43

121

(6H, s), 3.71-3.79 (4H, m), 3.86 (4H, q, J = 7.6 Hz), 4.02 (6H, s), 9.28 (2H, s), 9.90 (22H, s). 13C

NMR (125 MHz, CDCl3): δ 10.7, 11.5, 17.2, 18.2, 19.5, 19.8, 52.3, 94.4, 101.6, 109.9, 118.3,

135.0, 135.1, 135.4, 137.8, 139.8, 145.6, 157.2, 167.0. 1H NMR (500 MHz, DBU-CDCl3): δ -4.78

(1H, s), 9.22 (2H, s), 9.83 (2H, s); 1H NMR (500 MHz, 2 equiv. TFA-CDCl3): δ -5.41 (3H, s), -

1.5 (1H, v br), -1.42 (2H, s), 1.69 (6H, t, J = 7.7 Hz), 1.80 (6H, t, J = 7.7 Hz) 3.31 (6H, s), 3.74-

3.88 (6H, m), 3.88-3.96 (2H, m), 3.98 (6H, s), 9.49 (2H, s), 10.03 (2H, s). 13C NMR (125 MHz, 2

equiv. TFA-CDCl3): δ 11.1, 11.6, 16.7, 17.3, 19.66, 19.72, 52.3, 95.0, 113.4, 115.0, 121.8, 135.4,

138.0, 138.8, 139.7, 140.3, 141.3, 142.5, 166.4. 1H NMR (500 MHz, excess TFA-CDCl3): δ -8.67

(1H, q, J = 7.6 Hz), -6.55 (3H, d, J = 7.6 Hz), 1.94 (6H, t, J = 7.6 Hz), 1.97 (6H, t, J = 7.7 Hz),

3.85 (6H, s), 4.30-4.41 (8H, m), 4.66 (6H, s), 11.16 (2H, s), 11.80 (2H, s). 13C NMR (125 MHz,

excess TFA-CDCl3): δ 11.0, 12.0, 16.6, 17.7, 20.7, 20.8, 31.9, 55.5, 105.5, 114.4, 143.3, 144.3,

144.7, 145.9, 146.3, 149.2, 149.7, 150.6, 155.3, 166.9. HRMS (ESI): calcd for C36H41N3O4 + H

580.3175, found 580.3153.

122

8,12,13,17-Tetraethyl-2,3-di(methoxycarbonyl)-7,18,21,22-tetramethyl-21-carbaporphyrin (121)

Carbaporphyrin 98 (12.0 mg, 0.021 mmol) was stirred with potassium carbonate (12 mg) and

excess of methyl iodide (10 drops) in acetone (12 mL) under reflux for 16 h. The mixture was

washed with water and evaporated under reduced pressure. The residue was run through a grade

III neutral alumina column, eluting with dichloromethane and hexane in a ratio of 60:40, and a

brown colored fraction was collected. The solvent was evaporated under reduced pressure and the

residue recrystallized from chloroform-methanol to yield 121 (8.7 mg, 0.015 mmol, 69%) as purple

crystals, mp 222-224 oC; UV-Vis (CH2Cl2): λmax (log ε) 316 (4.41), 376 (4.59), 441 (5.89), 545

(3.93), 589 (4.01), 736 nm (3.30). UV-Vis (CH2Cl2): λmax (log ε) 320 (4.34), 412 (4.77), 429 (sh,

4.73), 567 (3.85), 607 nm (sh, 3.76). 1H NMR (500 MHz, CDCl3): δ -5.27 (3H, s), -3.91 (3H, s),

1.39 (3H, t, J = 7.6 Hz), 1.70 (3H, t, J = 7.7 Hz), 1.73-1.78 (6H, 2 overlapping triplets, J = 7.6 Hz),

2.95 (3H, s), 3.32 (3H, s), 3.37-3.43 (2H, m), 3.55-3.66 (3H, m), 3.68-3.76 (3H, m), 4.02 (3H, s),

4.06 (3H, s), 8.90 (1H, s), 9.05 (1H, s), 9.52 (1H, s), 9.54 (1H, s) (5,20-H). 13C NMR (125 MHz,

CDCl3): δ 9.6, 11.4, 11.7, 16.2, 17.1, 18.0, 18.1, 19.3, 19.57, 19.61, 19.8, 31.7, 52.1, 52.3, 93.7,

98.0, 108.9, 109.5, 112.5, 116.4, 123.3, 128.1, 131.2, 132.5, 134.4, 135.7, 136.1, 137.5, 141.8,

144.3, 145.8, 146.4, 150.6, 154.8, 158.2, 166.5, 167.7. 1H NMR (500 MHz, TFA-CDCl3): δ -4.91

123

(3H, s), -4.18 (1H, br s), -4.02 (1H, br s), -3.54 (3H, s), 1.46 (3H, t, J = 7.7 Hz), 1.72 (3H, t, J =

7.7 Hz), 1-80-1.85 (6H, 2 overlapping triplets, J = 7.7 Hz), 3.03 (3H, s), 3.37 (3H, s), 3.46-3.53

(2H, m), 3.82 (2H, q, J = 7.7 Hz), 3.86-3.97 (4H, m), 4.15 (3H, s), 4.19 (3H, s), 9.45 (1H, s), 9.46

(1H, s), 9.86 (1H, s), 9.92 (1H, s). 13C NMR (125 MHz, 2 µL TFA-CDCl3): δ 10.9, 11.8, 12.2,

15.8, 16.4, 17.17, 17.23, 19.6, 19.7, 19.8, 31.1, 53.0, 53.1, 92.9, 96.3, 113.3, 114.5, 117.4, 123.1,

134.3, 135.5, 136.5, 138.6, 138.9, 139.4, 141.1, 141.88, 141.94, 142.5, 144.6, 145.2, 150.6, 152.6.

HRMS (ESI): calcd for C37H43N3O4 + H 594.3332, found 594.3326.

124

(8,12,13,17-Tetraethyl-2,3-di(methoxycarbonyl)-7,18,21-trimethyl-21-

carbaporphyrinato)palladium(II) (122)

Palladium(II) acetate (10 mg, 0.044 mmol) was added to a solution of C-methylcarbaporphyrin

120 (10.0 mg, 0.017 mmol) in acetonitrile (10 mL), and the solution was stirred under reflux for

30 min. The solution was cooled, diluted with chloroform, washed with water, and the organic

layer separated and then evaporated to dryness. The residue was chromatographed with grade III

neutral alumina, eluting with 30% dichloromethane-hexanes, and the product was collected as a

green band. The solvent was evaporated to dryness and the residue recrystallized from chloroform-

methanol to yield the palladium(II) complex (7.4 mg, 0.011mmol, 63%) as dark green crystals, mp

282-284 oC; UV-Vis (CHCl3): λmax (log ε) 306 (4.37), 410 (4.54), 460 (4.52), 602 (3.90), 639 nm

(sh, 3.81). 1H NMR (500 MHz, CDCl3): δ -4.12 (3H, s), 1.81 (6H, t, J = 7.7 Hz), 1.83 (6H, t, J =

7.7 Hz), 3.44 (6H, s), 3.86-3.92 (8H, 2 overlapping quartets), 4.30 (6H, s), 9.85 (2H, s), 10.81 (2H,

s). 13C NMR (125 MHz CDCl3): δ 11.9, 17.6, 18.4, 19.8, 20.1, 21.8, 36.4, 52.9, 101.7, 113.0,

131.3, 136.9, 143.4, 143.9, 145.5, 145.6, 149.2, 158.6, 167.6. HRMS (ESI): calcd for

C36H39N3O4Pd + H 683.1851, found 683.1955.

125

(8,12,13,17-Tetraethyl-2,3-di(methoxycarbonyl)-7,18,21-trimethyl-21-

carbaporphyrinato)nickel(II) (123)

Nickel(II) acetate tetrahydrate (10 mg, 0.057 mmol) was added to a solution of C-

methylcarbaporphyrin 120 (10.0 mg, 0.017 mmol) in dimethylformamide (10 mL), and the

solution was stirred under reflux for 30 min. The solution was cooled, diluted with chloroform,

washed with water, and the organic layer separated and then evaporated to dryness. The residue

was chromatographed with grade III neutral alumina, eluting with 60% dichloromethane-hexanes,

and the product was collected as a green band. The solvent was evaporated to dryness and the

residue recrystallized from chloroform-methanol to yield the nickel(II) complex (7.6 mg, 0.012

mmol, 71%) as dark green crystals, mp 267-269 oC; UV-Vis (CHCl3): λmax (log ε) 348 (4.51), 420

(4.72), 456 (sh, 4.38), 612 nm (3.96). 1H NMR (500 MHz, CDCl3): δ -3.89 (3H, s), 1.72 (6H, t, J

= 7.7 Hz), 1.77 (6H, t, J = 7.7 Hz), 3.31 (6H, s), 3.72-3.81 (8H, m), 4.26 (6H, s), 9.59 (2H, s),

10.56 (2H, s). 13C NMR (125 MHz, CDCl3): δ 11.8, 17.5, 18.2, 19.7, 19.8, 19.9, 26.7, 52.8, 99.8,

111.0, 131.3, 138.3, 144.3, 145.9, 146.1, 148.7, 150.9, 167.6. HRMS (ESI): calcd for

C36H39N3O4Ni + H 635.2294, found 635.2285.

126

CHAPTER III

SYNTHESIS OF A CARBACHLORIN

Introduction

Chlorins (2,3-dihydroporphyrins) are an important group of tetrapyrroles that provide the

parent structure for many of the chlorophylls and other biological pigments present in nature.2

Chlorins tend to have a green coloration in solution because of their strong absorptions near 650

nm in the long wavelength region of visible spectrum. It is this absorption which is in part

responsible for their potential application as photosensitizers in photodynamic therapy,29 and this

property has led to numerous investigations into their synthesis.53 However, there have been very

few studies on the preparation of analogous systems, such as carbachlorins, that may have equally

useful properties.14 It was in 1998 that the first synthesis of carbachlorins was carried out by Hayes

and Lash. For this synthesis, a “3 + 1” variant on the MacDonald condensation was used.15 Bicyclic

dialdehydes were found to condense with a tripyrrane in the presence of TFA to give, following

oxidation with DDQ, carbachlorins 130 in moderate yields. These chlorin analogues were found

to retain the aromatic properties as the proton NMR spectra showed the inner CH protons upfield

around -7 ppm.35 The UV-vis spectra were also very porphyrin-like, showing a Soret band at 404

nm and three smaller absorptions between 490 and 650 nm. The band at 650 nm was of medium

intensity and resembles the characteristic absorptions of tetrapyrrolic chlorins.15

127

In 2014, Li and Lash, in their attempts to further explore carbachlorin chemistry,

synthesized a carbachlorin 131 using 3-ethoxymethylenecyclopentene-1-carbaldehyde. This was

condensed with a tripyrrane, and following oxidation with aqueous ferric chloride solutions,

generated moderate yields of a carbachlorin 131. Also, oxidation of the carbachlorin 131 with

DDQ gave the first example of a carbaporphyrin 132 that was unsubstituted on the carbocyclic

ring.55

However, compared to studies on other core modified porphyrins, for instance

carbaporphyrins, our knowledge of carbachlorins is still very limited. Questions such as whether

these core-modified chlorins (carbachlorins) share similar properties with core-modified

128

porphyrins (carbaporphyrins) is still open. Therefore, to further investigate the chemistry of

carbachlorins, it was necessary to develop an improved synthesis of a carbachlorin.

In the previous chapter, the synthesis of a carbaporphyrin with two methyl ester groups on

adjacent carbons was performed. It was interesting to note that carbachlorin 117 was also formed

as the minor fraction in this synthesis. However, this compound was prone to oxidation and

carbaporphyrin was always the major product in these reactions. In the case of carbachlorins 130,

the presence of a fused five-membered ring appeared to inhibit further oxidation. We speculated

that carbachlorins could be generated as the major products if the site of oxidation was blocked by

geminal substitution. Carbachlorin 133 was an appealing target molecule, where the methyl esters

were placed on the same carbon rather than at vicinal positions. It was anticipated that the synthesis

and study of carbachlorin 133 would increase our understanding of the chlorin system and could

allow the formation of new porphyrinoids and their metal complexes.

129

Results and Discussion

A retrosynthetic analysis suggested that carbachlorin 133 could be obtained from

cyclopentane dialdehyde 134 and the known tripyrrane 62. A straightforward acid catalyzed ‘3+1’

condensation of dialdehyde 134 with tripyrrane 62 in the presence of trifluoroacetic acid, followed

by oxidation with DDQ, would afford the targeted carbachlorin (Scheme 66).

Scheme 66 Retrosynthetic Analysis of Carbachlorin 133

In order to accomplish this goal, a synthesis of dialdehyde 134 had to be developed. This

was accomplished using a similar strategy to the one described in Chapter 2. Initially, norbornene

diester 135 was required and this could be generated using the Diels-Alder reaction. This required

dienophile 136, which was prepared by adapting a literature procedure reported by Harada and

coworkers.63 Copper(II) acetate was added to a solution of glacial acetic acid, paraformaldehyde

and dimethyl malonate. After stirring at 100 °C for 3 h, the reaction mixture was cooled to room

temperature and the excess acetic acid was removed under reduced pressure. The residue was later

distilled under reduced pressure to afford the colorless distillate 136 (Scheme 67). This distillate

was immediately diluted with CH2Cl2 and used for the following Diels-Alder reaction.57

130

Scheme 67 Formation of 136

In order to form the bicyclic adduct 135, freshly cracked 1,3-cyclopentadiene was added

dropwise to the solution of dimethyl 2-methylenemalonate 136 dissolved in CH2Cl2 whilst stirring.

The temperature of the reaction was controlled with a water bath. The reaction mixture was stirred

at room temperature for almost 60 min and purified by column chromatography on silica gel using

hexane/EtOAc (19:1, v/v) as the eluent to give bicyclic adduct 135 as a colorless oil (Scheme 68).57

Scheme 68 Formation of 135

Once the norbornene was generated, the next step was to oxidize the compound to dialcohol

137. Several different procedures were attempted to carry out the transformation, and it was

observed that the best results were obtained using KMnO4 and NaOH for oxidation. For this

reaction, a solution of potassium permanganate and sodium hydroxide in water was added

dropwise to a solution of norbornene in tert-butyl alcohol and water while maintaining the

131

temperature at 0° C. After the addition was completed, the reaction mixture was stirred for 40

minutes and quenched with a saturated aqueous solution of sodium metabisulfite until the solution

turned colorless. The mixture was filtered and the tert-butyl alcohol was removed from the filtrate

in vacuo. The remaining solution was extracted with ethyl acetate which was filtered and

concentrated in vacuo to yield dialcohol 137 as white crystalline solid (Scheme 69).59 Compound

137 had not been prepared previously and was synthesized for the first time.

Scheme 69 Formation of Dialdehyde 134

Diol 137 could be oxidatively ring opened using potassium periodate as an oxidizing agent

to give dialdehyde 134. Dialdehyde 134 was immediately condensed with tripyrrane dicarboxylic

acid 62 in the presence of TFA/CH2Cl2 and the reaction mixture stirred overnight under the

atmosphere of nitrogen. After stirring overnight, oxidation with DDQ in refluxing toluene was

performed and carbachlorin 133 was isolated as the major product (Scheme 70).35

132

Scheme 70 Preparation of Carbachlorin 133

Carbachlorin 133, like other chlorin analogues, proved to be aromatic in the free base form.

The proton NMR spectrum for 133 (Figure 41) in CDCl3 showed the presence of a strong

diamagnetic ring current. The internal CH proton was found at -7.08 ppm, while the meso-protons

were shifted downfield to 9.86, 9.84, 9.58 and 9.25 ppm. As there were four signals for the four

meso-protons, the molecule must be asymmetrical. The UV-Vis spectrum of carbachlorin 133,

shows a strong Soret band at 398 nm along with small Q bands at 491, 588 and 646 nm. The

slightly stronger absorption at 646 nm is characteristic of chlorin-type structures (Figure 42).

133

Figure 41 500 MHz Proton NMR Spectrum of Carbachlorin 133 in CDCl3

s

-7-6-5-4-3-2-110 9 8 7 6 5 4 3 2 1 0 ppm

9.49.69.8 ppm

-7.1 -7.2 ppm

3.83.94.0 ppm

134

Figure 42 UV-Vis Spectrum of Carbachlorin 133

When carbachlorin 133 was exposed to a trace amount of TFA, a deep purple solution

corresponding to monocation 133H+ (Scheme 71) was generated. The proton NMR spectrum for

133H+ (Figure 43) showed that the addition of acid to the carbachlorin system had slightly

increased its aromatic nature, as the internal CH was shifted upfield to -7.10 ppm, while the meso-

protons were shifted further downfield to 10.23, 10.22, 9.96 and 9.71 ppm (Figure 43). The

appearance of broad singlets at -4.95 ppm (2H) and -5.96 ppm (1H) confirmed the presence of an

additional NH due to protonation. Unlike the related carbaporphyrins, further addition of TFA to

0

5

10

15

20

25

300 350 400 450 500 550 600 650 700

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

135

133H+ did not give a diprotonated species, as aromatic character would be disrupted if protonation

occurred at the inner carbon.

Scheme 71 Protonation of 133

-7-6-5-4-3-2-112 11 10 9 8 7 6 5 4 3 2 1 0 ppm

-5.8 -6.0 -6.2 -6.4 -6.6 -6.8 -7.0 ppm

Figure 43 500 MHz Proton NMR Spectrum of Carbachlorin 133H+ in CDCl3

136

In order to further understand the reactivity of carbachlorin 133, a few more experiments

were performed to observe how the exchange of protons takes place with addition of acid to 133.

As expected, when d-TFA was added to a solution of 133 in CDCl3 the resulting monocation did

not show any exchange of the inner CH proton, in contrast to carbaporphyrins which exhibit rapid

deuterium exchange. However, exchange of the meso-protons was evident. After 20 minutes, the

intensities for two of the meso-resonances were reduced by 15%, while the remaining two peaks

were reduced by more than 50%. After several hours, proton-deuterium exchange was virtually

100% for all four meso-positions (Figure 44). The results imply that meso-substituted monocations

such as 133a-dH+ are in equilibria with 133 and 133H+ (Scheme 72). It is also possible that related

dications could also be involved. It is noteworthy that the exchange rate for chlorin 133 is far

greater than the exchange rates noted for carbaporphyrins such as 98.

137

Scheme 72 Proposed Equilibrium of Different Cationic Species for Carbachlorin 133

138

Figure 44 Proton NMR spectra of Carbachlorin 133 in CDCl3 with Two Drops of d-TFA after

20 min (upper spectrum) and 7 hours (lower spectrum).

Metalation of carbachlorin 133 with silver acetate was also briefly investigated. A solution

of carbachlorin 133 in CH2Cl2 was added to a suspension of silver(I) acetate in methanol, and the

resulting mixture was stirred overnight at room temperature. After purification by column

chromatography, a red fraction was collected and recrystallized from chloroform and methanol to

yield organometallic complex 138 as a dark red powder in 62% yield (Scheme 73). Silver complex

139

138 retains aromatic characteristics as indicated by proton NMR spectrum (Figure 45). The UV-

Vis spectrum of 138 (Figure 46) has a strong absorptions at 386 and 410 nm in the Soret region,

along with three absorptions at 490, 522 and 546 nm. Moreover, the stronger band at 592 nm is

reminiscent of chlorin systems.

Scheme 73 Formation of Silver(III) Chlorin 138

140

-8-7-6-5-4-3-2-113 12 11 10 9 8 7 6 5 4 3 2 1 0 ppm

9.39.49.59.69.79.89.910.010.1 ppm

Figure 45 500 MHz Proton NMR spectrum of 138 in CDCl3

141

Figure 46 UV-Vis Spectrum of Silver(III) Chlorin 138 in Dichloromethane

Conclusions

A cyclopentane dialdehyde was prepared in three steps, starting with the formation of

dimethyl 2-methylenemalonate 135 from paraformaldehyde and dimethyl malonate. This

dienophile then underwent Diels-Alder chemistry with cyclopentadiene to generate the bicyclic

compound 136. Norbornene 136 was then oxidized in two steps to generate dialdehyde 134.

Although difficulties were initially encountered in preparing dialdehyde precursor 134, careful

adjustments of the reaction conditions allowed the five-memebered ring dialdehyde to be isolated

in good yields. The ‘3+1’ condensation of dialdehyde 134 and tripyrrane dicarboxylic acid 62 led

to the formation of carbachlorin 133 in 30% yield. This newly synthesized carbachlorin system

0

2

4

6

8

10

12

14

16

300 350 400 450 500 550 600 650 700

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

142

proved to be highly aromatic. Protonation of this system can readily occur, forming an aromatic

monocation.

Reaction between carbachlorin 133 and silver(I) acetate gave a silver(III) organometallic

complex. The straightforward synthesis of carbachlorin 133 will allow further studies on this

system to be conducted. It is anticipated that further coordination chemistry could be performed

on this system as the yields are quite good.

Experimental

All chemicals were purchased from Sigma Aldrich or Acros Organic. Grade III neutral alumina

and silica gel were used to perform column chromatography. The 1H and 13C-NMR spectra were

obtained either using Bruker Avance III 500 MHz NMR spectrometer or Bukner 400 MHz NMR

spectrometer at 302K. Chemical shifts were recorded, relative to CDCl3 (residual chloroform at δ

7.26 ppm) in proton NMR spectra and the CDCl3 triplet at δ 77.23 ppm in 13C-NMR spectra, in

parts per million (ppm). The UV-vis spectra for the compounds were collected by Cary 100 Bio

spectrophotometer. Melting point were gathered using a Mel-Temp apparatus. Mass spectrometry

data was obtained from the Mass Spectral Laboratory, School of Chemical Sciences, University

of Illinois, Urbana-Champaign.

143

Dimethyl 2-methylenemalonate (136)63

Paraformaldehyde (20 g, 133 mmol), dimethyl malonate (50 mL, 87.5 mmol), and

copper(II) acetate (3.25g, 1.64 mmol) were added to glacial acetic acid (100 ml). The reaction

mixture was stirred at 100 °C for 3 h, then cooled to room temperature, and the acetic acid removed

under reduced pressure. The residue was distilled under reduced pressure using a vacuum pump

and a colorless fraction was obtained between 90 oC to 97 oC (the external temperature was held

between 152 oC to 160 oC) corresponding to 137 (3.6 g, 0.025mmol, 28%). This product was

immediately diluted with CH2Cl2 and the solution used for the following Diels-Alder reaction. 1H

NMR (500 MHz, CDCl3): δ 6.60 (2H, s), 4.02 (6H, s).

144

5,5-Dimethoxycarbonyl-2-norbornene (135)57

A solution of dimethyl 2-methylenemalonate 137 in dichloromethane (1.37 g, 0.021mmol) was

added slowly to freshly cracked 1,3-cyclopentadiene (3.00 g, 0.021 mol) whilst stirring and

maintaining temperature at 25oC with the aid of water bath. The reaction mixture was stirred at room

temperature for 60 min and purified by chromatography on silica gel using hexane/EtOAc (19:1) to

give the Diels-Alder adduct as a colorless oil (2.97 g, 0.0141 mmol, 80%). 1H NMR (500MHz,

CDCl3): δ 1.48-1.52 (1H, m), 1.65 (1H, br d, J = 8.9 Hz), 1.99 (1H, dd, J = 12.4, 2.9 Hz), 2.10

(1H, dd, J = 12.4, 3.6 Hz), 2.90 (1H, br), 3.37-3.39 (1H, br m), 3.65 (3H, s), 3.72 (3H, s), 5.97

(1H, dd, J = 5.6, 2.9 Hz), 6.24 (1H, dd, J = 5.6, 3.1 Hz). 13C NMR (125 MHz, CDCl3): δ 36.1,

42.2, 48.9, 50.0, 52.5, 52.9, 60.4, 133.7, 140.0, 171.6, 173.2. HRMS (EI): calcd for C11H14O4

210.0892, found 210.0896.

145

Dimethyl 5,6-dihydroxybicyclo[2.2.1]heptane-2,2-dicarboxlate (137)

To a solution of norbornene 137 (10.80 g, 53.1 mmol) in tert-butyl alcohol (200 mL) and water

(50 mL) at 0 °C was added dropwise a solution of potassium permanganate (12.0 g, 76.0 mmol)

and sodium hydroxide (2.61 g, 65.3 mmol) in water (250 mL). After the addition was complete,

the reaction mixture was stirred for 20 minutes and then quenched using a saturated aqueous

solution of sodium metabisulphite (Na2S2O5) until the solution turned colorless. The mixture was

filtered and the tert-butyl alcohol was removed from the filtrate in vacuo. The solution was

extracted with ethyl acetate (4 × 100 mL) and the combined organic layers dried over MgSO4,

filtered and concentrated in vacuo to furnish the diol 138 (6.23 g, 0.025 mmol, 61%) as a white

crystalline solid, mp 86-88 oC. 1H NMR (500 MHz, CDCl3): δ 1.43 (1H, dq, J = 11.0, 1.6 Hz),

1.88 (1H, dq, J = 11.0, 1.6 Hz), 2.49-2.50 (1H, m), 2.54-2.56 (1H, m), 2.72 (1H, dd, J = 5.6, 1.5

Hz), 2.86 (1H, br), 2.96 (1H, br), 3.20 (1H, dd, J = 5.6, 4.7 Hz), 3.69 (3H, s), 3.71 (3H, s), 3.75-

3.77 (1H, m), 3.85-3.87 (1H, m). 13C NMR (125 MHz, CDCl3): δ 31.9, 45.0, 46.4, 46.7, 48.5,

52.4, 52.5, 70.5, 73.6, 173.2, 174.3. HRMS (EI): calcd for C11H16O6 + H 245.1025, found

245.1017.

146

Dimethyl 3,5-diformylcyclopentane-1,1-dicarboxylate (134)

The foregoing diol 138 (3.05 g, 0.021 mole) was mixed with water (5.0 ml), ether (10 ml) and

potassium periodate (4.16 g, 0.018 mol) and the reaction stirred at room temperature for 1 hr. The

excess potassium periodate was filtered off and the two phases were separated. The water layer

was back extracted twice with ether. The ether layers were combined, dried over sodium sulfate

and the ether removed under reduced pressure to form the dialdehyde (1.67 g, 0.021 mmol, 60%)

as a clear oil. The unstable dialdehyde was not further purified but used immediately to prepare

carbachlorin. 1H NMR (500 MHz, CDCl3): 2.17-2.25 (1H, m), 2.47 (1H, dt, J = 13.8, 5.9 Hz),

3.05-3.11 (1H, m), 3.27-3.32 (1H, m), 3.57-3.66 (2H, m), 3.70 (3H, s), 3.76 (3H, s), 9.65 (1H, d,

J = 0.8 Hz), 9.68 (1H, d, J = 0.9 Hz). HRMS (EI): calcd for C11H14O6 + H 243.0869, found

243.0870.

147

8,12,13,17-Tetraethyl-2,3-dimethoxycarbonyl-7,18-dimethyl-21-carbaporphyrin (133)

Tripyrrane dicarboxylic acid 62 (100 mg, 0.220 mmol) was stirred with TFA (1 ml) under an

atmosphere of nitrogen for 2 minutes in a 250 ml round bottom flask. Dichloromethane (99 ml)

was added, followed immediately by dialdehyde 140 (55 mg, 0.23 mmol) and the mixture was

stirred under nitrogen overnight. The mixture was neutralized by the addition of triethylamine and

the solvent removed under reduced pressure. The residue was taken up in toluene (50 ml), DDQ

(100 mg, 0.44 mmol) was added, and the resulting mixture was stirred under reflux for 1 hr. The

solvent was removed under reduced pressure, then the residue was dissolved in CH2Cl2 and washed

with water. The solution was concentrated under reduced pressure and the residue

chromatographed on grade III neutral alumina column eluting with CH2Cl2 and hexane in a ratio

of 60:40.The product was collected as a dark green fraction. Recrystallization from chloroform

and methanol gave the carbachlorin (37 mg, 0.065 mmol, 30%) as purple crystals, mp 228-230 oC.

UV-Vis (1% Et3N-CH2Cl2): λmax (log ε) 398 (5.36), 491 (4.23), 588 (3.45), 646 nm (4.31). UV-

Vis (5% TFA-CH2Cl2): λmax (log ε) 406 (5.30), 423 (5.19), 527 (3.86), 543 (3.86), 583 (3.93), 635

nm (3.99). 1H NMR (500 MHz, CDCl3): δ -7.08 (1H, s), 1.80-1.84 (6H, 2 overlapping triplets, J =

7.7 Hz), 1.85-1.88 (6H, 2 overlapping triplets, J = 7.7 Hz), 3.50 (3H, s), 3.54 (3H, s), 3.90 (6H, s),

148

3.96-4.01 (4H, 2 overlapping quartets), 4.04-4.09 (4H, 2 overlapping quartets), 5.43 (2H, s), 9.25

(1H, s), 9.58 (1H, s), 9.84 (1H, s), 9.86 (1H, s). 13C NMR (125 MHz, CDCl3): δ 11.4, 11.6, 17.6,

18.8, 19.8, 20.2, 45.3, 53.5, 96.2, 96.5, 97.7, 99.3, 120.7, 129.7, 130.2, 132.95, 133.05, 135.5,

135.8, 137.8, 137.9, 139.8, 140.6, 143.3, 143.4, 150.5, 172.2. 1H NMR (500 MHz, 5% TFA-

CDCl3): δ -7.10 (1H, s), -5.96 (1H, br s), -4.95 (2H, br s), 1.67-1.72 (6H, 2 overlapping triplets),

1.83-1.87 (2 overlapping triplets), 3.49 (3H, s), 3.51 (3H, s), 3.98 (6H, s), 4.09 (4H, q, J = 7.7 Hz),

4.16 (4H, q, J = 7.6 Hz), 9.71 (1H, s), 9.96 (1H, s), 10.226 (1H, s), 10.231 (1H, s). 13C NMR (125

MHz, 5% TFA-CDCl3): δ 11.44, 11.47, 16.6, 17.6, 20.1, 20.2, 45.7, 55.1, 68.4, 94.6, 94.7, 103.7,

103.9, 128.6, 134.5, 134.8, 134.9, 135.1, 135.6, 135.9, 141.4, 142.0, 142.1, 142.7, 142.9, 143.0,

145.0, 173.7. HRMS (ESI): calcd for C35H41N3O4 + H 568.3175, found 568.3177.

149

(8,12,13,17-Tetraethyl-2,2-dimethoxycarbonyl-7,18-dimethyl-21-carbaporphyrinato)silver(III)

(138)

A solution of carbachlorin 133 (10.0 mg, 0.017 mmol) in dichloromethane (10 mL) was added to

silver(I) acetate (12 mg, 0.019 mmol) dissolved in methanol (2.5 mL), and the mixture was stirred

at room temperature overnight. The mixture was washed with water, and the solvent was removed

under reduced pressure. The residue was chromatographed on grade III neutral alumina eluting

with dichloromethane and hexane in a ratio of 60:40. A deep red fraction was collected and

recrystallized from chloroform-methanol to give the silver(III) complex (7.3 mg, 0.011mmol,

62%) as a red powder, mp 266-268 oC. UV-Vis (CH2Cl2): λmax (log ε) 386 (sh, 4.49), 410 (5.15),

490 (3.96), 522 (3.82), 546 (3.58), 592 nm (4.43). 1H NMR (500 MHz, CDCl3): 1.81 (3H, t, J =

7.7 Hz), 1.83 (3H, t, J = 7.7 Hz), 1.88 (6H, t, J = 7.7 Hz), 3.47 (3H, s), 3.54 (3H, s), 3.93 (6H, s),

3.98-4.09 (8H, m), 9.36 (1H, s), 9.70 (1H, s), 9.93 (1H, s), 9.98 (1H, s) (10,15-H). 13C NMR (125

MHz CDCl3): δ 11.8, 12.0, 17.82, 17.83, 18.58, 18.59, 19.9, 20.48, 20.52, 45.3, 53.5, 67.5, 98.2,

98.5, 101.3, 103.0, 130.8, 133.4, 133.5, 134.2, 134.4, 136.0, 136.5, 138.7, 138.8, 140.3, 140.6,

172.4. HRMS (ESI): calcd for C35H38AgN3O4 + H 672.1992, found 672.2007.

150

CHAPTER IV

SUMMARY AND CONCLUSIONS

Carbaporphyrins and carbachlorins are interesting molecules showing strong diatropic

character and unusal coordination chemistry. In order to expand this important area, new routes to

carbaporphyrins and carbachlorins have been developed. In Chapter II, it was shown that a

carbaporphyrin could be synthesized from a cyclopentane dialdehyde in good yields and this

porphyrin analogue proved to be a versatile precursor for alkylation and coordination chemistry.

Initially, a Diels-Alder adduct was prepared from cyclopentadiene and dimethyl fumarate. This

was converted in two steps to the cyclopentane dialdehyde, and subsequent condensation with a

tripyrrane in the presence of TFA, followed by oxidation with DDQ in refluxing toluene, gave the

novel carbaporphyrin in 40% yield together with a low yield of a related carbachlorin. The

carbaporphyrin possesses macrocyclic aromaticity and afforded stable silver(III) and gold(III)

complexes. In addition, alkylation with methyl iodide and potassium carbonate gave a C-methyl

derivative and this underwent metalation with palladium(II) acetate and nickel(II) acetate to give

the related organometallic species. These reactions involved both the formation of an

organometallic species and an increase in aromaticity. On protonation, the new carbaporphyrin

also exhibited enhanced diatropic character. Mono- and diprotonated species were characterized,

and deuterium exchange studies also indicated that meso-protonated structures were present in

equilibrium. In addition to characterization by UV-Vis, proton NMR and carbon-13 NMR

spectroscopy, and mass spectrometry, the structures of the free base carbaporphyrin and two

metalated derivatives were confirmed by X-ray crystallography.

The ‘3+1’ variant of the MacDonald condensation has been widely used to prepare

151

porphyrin analogues but only one example of using aliphatic dialdehydes has previously been

reported for this synthetic strategy. Even in that case, only poor yields of porphyrinoid products

were obtained. In this work, aliphatic cyclopentane dialdehydes bearing two ester substituents

were shown to be superior precursors to carbaporphyrinoids. The efficient synthesis of a

carbaporphyrin with ester substituents lays the foundation for further investigation.

Future work entailing this carbaporphyrin could include reductions to generate dicarbinols

that could be converted into precursors for cycloaddition reactions. This strategy would provide

access to further modified structures.

In Chapter III, it was demonstrated that an equally efficient synthesis of a carbachlorin

could be facilitated using a cyclopentane dialdehyde with geminal diester substituents. The

required dialdehyde was obtained from the Diels-Alder adduct of dimethyl methylenemalonate

and cyclopentadiene. Condensation of the dialdehyde with a tripyrrane in the presence of

trifluoroacetic acid, followed by oxidation with DDQ, produced a novel carbachlorin diester in

30% yield. This macrocycle was found to react with silver(III) acetate forming a silver(III)

complex in good yields. The carbachlorin and its silver complex both exhibited strongly aromatic

characteristics. Addition of acid to the carbachlorin resulted in the formation of a monocation with

slightly enhanced diatropic character.

The methodology reported in the thesis provides the most convenient route available for

preparing carbaporphyrins and carbachlorins. The availability of the new porphyrin analogues will

allow further investigations into the reactivity of these systems. Furthermore, the presence of

absorption bands near 650 nm for carbachlorins may allow this system to be utilized as a

photosenitizers in photodynamic therapy.

152

REFERENCES

1. Porphyrins and Metalloporphyrins; Smith, K. M., ed.; Elsevier, 1975.

2. Milgrom, L. R. The Colours of Life: An Introduction to the Chemistry of Porphyrins and

Related Compounds; Oxford University Press, 1997.

3. Kekulé, A. Bull de la Societe Chimique de Paris, 1865, 3, 98–110.

4. Loudon, G. M. Organic Chemistry; Roberts and Company, 2009.

5. Hofmann, A. W. Proc. R. Soc. Lond. 1856, 8, 1–3.

6. Sondheimer, F.; Wolovsky, R.; Amiel, Y. J Am. Chem. Soc. 1962, 84, 274–284.

7. Lash, T. D. in The Porphyrin Handbook: Heteroporphyrins, Expanded Porphyrins and

Related Macrocycles; Kadish, K. M.; Smith, K. M.; Guilard, R. Academic Press, 2000;

Vol. 2, pp 125–199.

8. Soret, J. L. Compt. Rend. 1883, 97, 1267.

9. Kuster, W.; Deihle, P. Z. Physiol. Chem. 1912, 82.

10. Fischer, H.; Zeile, K. Justus Liebigs Ann. Der Chem. 1929, 468, 98–116.

11. Smith, K. M. in The Porphyrin Handbook: Synthesis and Organic Chemistry; Kadish, K.

M.; Smith, K. M.; Guilard, R Academic Press, 2000; Vol. 1, pp 2–43.

12. Rothemund, P. J. Am. Chem. Soc. 1936, 58, 625–627.

13. Adler, A. D.; Longo, F. R.; Finarelli, J. D.; Goldmacher, J.; Assour, J.; Korsakoff, L. J.

Org. Chem. 1967, 32, 476–477.

14. Lindsey, J. S.; Schreiman, I. C.; Hsu, H. C.; Kearney, P. C.; Marguerettaz, A. M. J. Org.

Chem. 1987, 52, 827–836.

15. Arsenault, G. P.; Bullock, E.; MacDonald, S. F. J. Am. Chem. Soc. 1960, 82, 4384–4389.

16. Broadhurst, M. J.; Grigg, R.; Johnson, A. W. J. Chem. Soc. 1971, 3681–3690.

17. Cavaleiro, J. A. S.; Gonsalves, A. M. d’A R.; Kenner, G. W.; Smith, K. M. J. Chem. Soc

Perkin Trans. 1973, 1, 2471.

18. Sessler, J. L.; Johnson, M. R.; Lynch, V. J. Org. Chem. 1987, 52, 4394–4397.

153

19. Lash, T. D.; Toney, A. M.; Castans, K. M.; Ferrence, G. M. J. Org. Chem. 2013, 78,

9143–9152.

20. Graham, S. R.; Colby, D. A.; Lash, T. D. Angew. Chem. Int. Ed. 2002, 41, 1371–1374.

21. Lash, T. D.; Chaney, S. T.; Richter, D. T. J. Org. Chem. 1998, 63, 9076–9088.

22. Sessler, J. L.; Cyr, M.; Burrell, A, K. Tetrahedron, 1992, 48, 9661–9672.

23. Lash, T. D. Eur. J. Org. Chem. 2007, 33, 5461–5481.

24. Muckey, M. A.; Szczepura, L. F.; Ferrence, G. M.; Lash, T. D. Inorg. Chem. 2002, 41,

4840 – 4842.

25. Xue, Z.-L.; Shen, Z.; Mack, J.; Kuzuhara, D.; Yamada, H.; Okujima, T.; Ono, N.;

You, X.-Z.; Kobayashi, N. J. Am. Chem. Soc. 2008, 130, 16478–16479.

26. Broadhurst, M. J.; Grigg, R.; Johnson, A. W. J. Chem. Soc. Perkin Trans. 1 1972, 2111.

27. Sessler, J. L.; Murai, T.; Lynch, V.; Cyr, M. J. Am. Chem. Soc. 1988, 110, 5586–5588.

28. Amao, T.; Okura, I. J. Porphyrins Phthalocyanines 2009, 13, 1111–1122.

29. Malinski, T. in The Porphyrin Handbook: Applications Past, Present, and Future;

Kadish, K. M, Smith, K. M.; Guilard, R. Academic Press, 2000; Vol.6, pp 232–255.

30. Collman, J. P.; Brauman, J. I.; Meunier, B.; Raybuck, S. A.; Kodadek, T. Proc Natl. Acad

Sci. 1984, 81, 3245–3248.

31. Anding, B. J.; Ellern, A.; Woo, L. K. Organometallics, 2012, 31, 3628–3635.

32. Kostas, I. D.; Coutsolelos, A G; Charalambidis, G; Skondra, A. Tetrahedron Lett. 2007,

48, 6688–6691.

33. Rambo, B. M.; Sessler, J. L. Chem. –Eur. J. 2011, 17, 4946–4959.

34. Groves, J. T.; Myers, R. S. J. Am. Chem. Soc. 1983, 105, 5791–5796.

35. Lash, T. D. Chem. Rev. 2017, 117, 2313–2446.

36. Toganoh, M.; Furuta, H. Chem. Commun. 2012, 48, 937−954.

37. Bergman, K. M.; Ferrence, G. M.; Lash, T. D. J. Org. Chem. 2004, 69, 7888−7897.

38. Lash, T. D. Acc. Chem. Res. 2016, 49, 471−482.

154

39. Lash, T. D. Org. Biomol. Chem. 2015, 13, 7846−7878.

40. Lash, T. D.; Hayes, M. J. Angew. Chem., Int. Ed. Engl. 1997, 36, 840−842.

41. AbuSalim, D. I.; Lash, T. D. J. Org. Chem. 2013, 78, 11535−11548.

42. Colby, D. A.; Lash, T. D. Chem. - Eur. J. 2002, 8, 5397−5402.

43. Lash, T. D.; Colby, D. A.; Ferrence, G. M. Eur. J. Org. Chem. 2003, 2003, 4533−4548.

44. Lash, T. D.; Hayes, M. J. Angew. Chem., Int. Ed. Engl. 1997, 36, 840−842.

45. Lash, T. D.; Hayes, M. J.; Spence, J. D.; Muckey, M. A.; Ferrence, G. M.; Szczepura, L.

F. J. Org. Chem. 2002, 67, 4860−4874.

46. Muckey, M. A.; Szczepura, L. F.; Ferrence, G. M.; Lash, T. D. Inorg. Chem. 2002, 41,

4840−4842.

47. Lash, T. D.; Colby, D. A.; Szczepura, L. F. Inorg. Chem. 2004, 43, 5258−5267.

48. Adiraju, V. A. K.; Ferrence, G. M.; Lash, T. D. Dalton Trans. 2016, 45, 13691−13694.

49. Lash, T. D. Org. Lett. 2011, 13, 4632−4635.

50. Hayes, M. J.; Spence, J. D.; Lash, T. D. Chem. Commun. 1998, 34, 2409−2410.

51. Lash, T. D.; Muckey, M. A.; Hayes, M. J.; Liu, D.; Spence, J. D.; Ferrence, G. M. J. Org.

Chem. 2003, 68, 8558−8570.

52. Lash, T. D.; Romanic, J. L.; Hayes, M. J.; Spence, J. D. Chem. Commun. 1999, 35,

819−820.

53. Taniguchi, M.; Lindsey, J. S. Chem. Rev. 2017, 117, 344–535.

54. Hayes, M. J.; Lash, T. D. Chem. - Eur. J. 1998, 4, 508−511.

55. Daming, Li.; Lash, T. D. J. Org. Chem. 2014, 79, 7112–7121.

56. Barton, D. H. R.; Zard, S. Z. J. Chem. Soc. 1985, 16, 1098.

155

57. Atkins, A. J.; Fairlamb, I. J. S.; Ward, J. S.; Lynam, J. M. Organometallics 2012, 31,

5896–5902.

58. Göksu, S.; Altundas, R.; Sütbeyaz, Y. Syn. Comm. 2000, 9, 1615–1621.

59. Donohoe, T.; Jahanshahi, A.; Tucker, M. J.; Bhatti, F. L.; Roslan, I. A.; Kabeshov, M.;

Wrigley, G. Chem. Commun. 2011, 47, 5849–5851.

60. Henderson, L. C.; Li, J.; Nation, R. L.; Velkov, T.; Pfeffer, F. M.; Chem. Commun. 2010,

46, 3197–3199.

61. Hickey, S. M.; Ashton, T. D.; White, J. M.; Li, J.; Nation. R. L.; Yu, H. Y.; Elliott, A. G.;

Butler, M. S.; Huang, J. X.; Cooper, M. A.; Pfeffer, F. M. RSC Adv. 2015, 5, 28582–

28596.

62. Beare, K. D.; Yuen, A. K. L.; Masters, A. F.; Maschmeyer, T.; McErlean, C. S. P. Chem.

Commun. 2013, 49, 8347–8349.

63. Harada, M.; Asaba, K. N.; Iwai, M.; Kogure, N.; Kitajima, M.; Org. Lett., 2012, 14,

5800l.

156

APPENDIX A

SELECTED NMR SPECTRA

Figure A – 1: 400 MHz 1H NMR Spectrum of 114 in CDCl3

157

Figure A – 2: 100 MHz carbon-13 NMR spectrum of 114 in CDCl3.

158

12345678910 ppm

1.4

88

1.4

92

1.4

95

1.5

10

1.5

14

1.5

17

1.6

34

1.6

56

1.9

76

1.9

83

2.0

07

2.0

14

2.0

85

2.0

94

2.1

16

2.1

25

2.9

01

3.3

85

3.6

51

3.7

19

5.9

57

5.9

65

5.9

71

5.9

79

6.2

34

6.2

42

6.2

48

6.2

56

7.2

60

1.0

5

1.1

7

1.0

4

1.0

5

1.0

2

1.0

0

3.0

6

3.1

0

1.0

0

1.0

1

1.901.952.002.052.102.152.20 ppm

6.006.056.106.156.206.256.30 ppm

Figure A – 3: 400 MHz 1H NMR Spectrum of 135 in CDCl3

159

Figure A – 4: 100 MHz carbon-13 NMR spectrum of 135 in CDCl3.

160

Figure A – 5: 400 MHz 1H NMR Spectrum of 115 in CDCl3

161

Figure A – 6: 100 MHz carbon-13 NMR spectrum of 115 in CDCl3.

162

Figure A – 7: 400 MHz 1H NMR Spectrum of 137 in CDCl3

163

Figure A – 8: 100 MHz carbon-13 NMR spectrum of 137 in CDCl3.

164

Figure A – 9: 500 MHz 1H NMR Spectrum of Carbaporphyrin 98 in CDCl3

165

Figure A – 10: 1H-1H COSY NMR spectrum of carbaporphyrin diester 98 in CDCl3

Figure A – 11: HSQC NMR spectrum of carbaporphyrin diester 98 in CDCl3.

166

Figure A – 12: Selected nOe difference proton NMR spectra of carbaporphyrin diester 98 in

CDCl3.

2.02.53.03.54.04.55.05.56.06.57.07.58.08.59.09.510.010.5 ppm

167

Figure A – 13: DEPT-135 NMR spectrum of carbaporphyrin diester 98 in CDCl3.

Figure A – 14: 125 MHz carbon-13 NMR spectrum of carbaporphyrin diester 98 in CDCl3.

168

Figure A – 15: 500 MHz 1H NMR Spectrum of C-methylated Carbaporphyrin 120 in CDCl3

169

Figure A – 16: Selected nOe difference proton NMR spectra of 120 in CDCl3

2.02.53.03.54.04.55.05.56.06.57.07.58.08.59.09.510.0 ppm

170

Figure A – 17: 1H-1H COSY NMR spectrum of 120 in CDCl3

Figure A – 18: HSQC NMR spectrum of 120 in CDCl3

171

Figure A – 19: DEPT-135 NMR spectrum of 120 in CDCl3.

Figure A –20: 125 MHz carbon-13 NMR spectrum of 120 in CDCl3.

172

Figure A – 21: 500 MHz 1H NMR Spectrum of 121 in CDCl3.

173

Figure A – 22: Selected nOe difference proton NMR spectra of 121 in CDCl3

2.02.53.03.54.04.55.05.56.06.57.07.58.08.59.09.510.010.5 ppm

174

Figure A – 23: 1H-1H COSY NMR spectrum of 121 in CDCl3

Figure A – 24: HSQC NMR spectrum of 121 in CDCl3

175

Figure A – 25: DEPT-135 NMR spectrum of 121 in CDCl3

Figure A – 26: 125 MHz carbon-13 NMR spectrum of 121 in CDCl3

176

Figure A – 27: 500 MHz 1H NMR Spectrum of 118 in CDCl3

177

Figure A – 28: Selected nOe difference proton NMR spectra of 118 in CDCl3

10 9 8 7 6 5 4 3 2 1 ppm

178

Figure A – 29: 1H-1H COSY NMR spectrum of 118 in CDCl3

Figure A – 30: HSQC NMR spectrum of 118 in CDCl3

179

Figure A – 31: DEPT-135 NMR spectrum of 118 in CDCl3

Figure A – 32: 125 MHz carbon-13 NMR spectrum of 118 in CDCl3

160 150 140 130 120 110 100 90 80 70 60 50 40 30 20 10 ppm

1.2

3

11

.53

17

.30

18

.10

19

.39

19

.95

29

.92

31

.11

52

.61

76

.97

77

.23

77

.48

96

.13

11

0.1

6

11

8.3

2

13

1.0

9

13

4.5

2

13

4.5

5

13

4.5

8

13

7.1

6

13

8.3

1

13

9.6

4

13

9.6

6

14

0.9

0

16

7.4

5

18.018.519.019.520.0 ppm

180

Figure A – 33: 500 MHz 1H NMR Spectrum of 138 in CDCl3

181

Figure A – 34: Selected nOe difference proton NMR spectra of 138 in CDCl3

10 9 8 7 6 5 4 3 2 1 ppm

182

Figure A – 35: 1H-1H COSY NMR spectrum of 138 in CDCl3

Figure A – 36: HSQC NMR spectrum of 138 in CDCl3

183

Figure A – 37: DEPT-135 NMR spectrum of 138 in CDCl3

Figure A – 38: 125 MHz carbon-13 NMR spectrum of 138 in CDCl3

184

-7-6-5-4-3-2-114 13 12 11 10 9 8 7 6 5 4 3 2 1 0 ppm

-0.0

47

0.0

27

0.0

48

0.0

65

0.0

73

0.0

80

0.0

90

0.1

08

0.1

53

0.1

89

1.2

58

1.5

39

1.6

31

1.6

47

1.6

53

1.6

62

1.6

68

1.6

84

3.3

62

3.6

82

3.6

98

3.7

31

3.7

46

4.3

66

7.2

60

9.0

16

9.9

90

12.1

3

6.2

5

4.1

9

4.0

4

6.1

2

2.0

2

2.0

0

3.33.43.53.63.73.83.94.04.14.24.34.4 ppm

NN

N

Et

Me

MeO2C

MeO2C Me

Et

Et

Et

Au

Figure A – 39: 500 MHz 1H NMR Spectrum of 119 in CDCl3

185

Figure A – 40: Selected nOe difference proton NMR spectra of 119 in CDCl3

10 9 8 7 6 5 4 3 2 1 ppm

186

Figure A – 41: 1H-1H COSY NMR spectrum of 119 in CDCl3

Figure A – 42: HSQC NMR spectrum of 119 in CDCl3

187

Figure A – 43: DEPT-135 NMR spectrum of 119 in CDCl3

Figure A – 44: 125 MHz carbon-13 NMR spectrum of 119 in CDCl3

188

Figure A – 45: 500 MHz 1H NMR Spectrum of 133 in CDCl3

189

Figure A – 46: Selected nOe difference proton NMR spectra of 133 in CDCl3

10 9 8 7 6 5 4 3 2 1 ppm

190

Figure A – 47: 1H-1H COSY NMR spectrum of 133 in CDCl3

Figure A – 48: HSQC NMR spectrum of 133 in CDCl3

191

Figure A – 49: DEPT-135 NMR spectrum of 133 in CDCl3

Figure A – 50: 125 MHz carbon-13 NMR spectrum of 133 in CDCl3

192

Figure A – 51: 500 MHz 1H NMR Spectrum of 117 in CDCl3

193

Figure A – 52: Selected nOe difference proton NMR spectra of 117 in CDCl3

12345678910 ppm

194

Figure A – 53: 1H-1H COSY NMR spectrum of 117 in CDCl3

Figure A – 54: HSQC NMR spectrum of 117 in CDCl3

195

Figure A – 55: DEPT-135 NMR spectrum of 117 in CDCl3

Figure A – 56: 125 MHz carbon-13 NMR spectrum of 117 in CDCl3

196

Figure A – 57: 500 MHz 1H NMR Spectrum of 122 in CDCl3

197

Figure A – 58: Selected nOe difference proton NMR spectra of 122 in CDCl3

1234567891011 ppm

198

Figure A – 59: 1H-1H COSY NMR spectrum of 122 in CDCl3

Figure A – 60: HSQC NMR spectrum of 122 in CDCl3

199

Figure A – 61: DEPT-135 NMR spectrum of 122 in CDCl3

Figure A – 62: 125 MHz carbon-13 NMR spectrum of 122 in CDCl3

200

Figure A – 63: 500 MHz 1H NMR Spectrum of 123 in CDCl3

201

Figure A – 64: Selected nOe difference proton NMR spectra of 123 in CDCl3

10 9 8 7 6 5 4 3 2 1 ppm

202

Figure A – 65: 1H-1H COSY NMR spectrum of 123 in CDCl3

Figure A – 66: HSQC NMR spectrum of 123 in CDCl3

203

Figure A – 67: DEPT-135 NMR spectrum of 123 in CDCl3

Figure A – 68: 125 MHz carbon-13 NMR spectrum of 123 in CDCl3

204

Figure A – 69: 500 MHz 1H NMR Spectrum of 98H+ in TFA-CDCl3

205

Figure A – 70: Selected nOe difference proton NMR spectra of 98H+ in TFA-CDCl3

10 9 8 7 6 5 4 3 2 1 ppm

206

Figure A – 71: DEPT-135 NMR spectrum of 98H+ in TFA-CDCl3

Figure A – 72: HSQC NMR spectrum of 98H+ in TFA-CDCl3

207

Figure A – 73: DEPT-135 NMR spectrum of 98H+ in TFA-CDCl3

Figure A – 74: 125 MHz carbon-13 NMR spectrum of 98H+ in TFA-CDCl3

208

Figure A – 75: 500 MHz 1H NMR Spectrum of 98H22+ in TFA-CDCl3

209

Figure A – 76: 1H-1H COSY NMR spectrum of 98H22+ in TFA-CDCl3

Figure A – 77: HSQC NMR spectrum of 98H22+ in TFA-CDCl3

210

Figure A – 78: DEPT-135 NMR spectrum of 98H22+ in TFA-CDCl3

Figure A – 79: 125 MHz carbon-13 NMR spectrum of 98H22+ in TFA-CDCl3

211

Figure A – 80: 500 MHz 1H NMR Spectrum of 120H+ in TFA-CDCl3

212

Figure A – 81: Selected nOe difference proton NMR spectra of 120H+ in TFA-CDCl3

10 9 8 7 6 5 4 3 2 1 ppm

213

Figure A – 82: 1H-1H COSY NMR spectrum of 120H+ in TFA-CDCl3

Figure A – 83: HSQC NMR spectrum of 120H+ in TFA-CDCl3

214

Figure A – 84: 125 MHz carbon-13 NMR spectrum of 120H+in TFA-CDCl3

215

Figure A – 85: 500 MHz 1H NMR Spectrum of 120H22+ in TFA-CDCl3

216

Figure A – 86: 1H-1H COSY NMR spectrum of 120H22+ in TFA-CDCl3

Figure A – 87: HSQC NMR spectrum of 120H22+ in TFA-CDCl3

217

Figure A – 88: DEPT-135 NMR spectrum of 120H22+ in TFA-CDCl3

Figure A –89: 125 MHz carbon-13 NMR spectrum of 120H22+ in TFA-CDCl3

218

Figure A – 90: 500 MHz 1H NMR Spectrum of 133H+ in TFA-CDCl3

219

Figure A – 91: Selected nOe difference proton NMR spectra of 133H+ in TFA-CDCl3

11 10 9 8 7 6 5 4 3 2 1 ppm

220

Figure A – 92: 1H-1H COSY NMR spectrum of 133H+ in TFA-CDCl3

Figure A – 93: HSQC NMR spectrum of 133H+ in TFA-CDCl3

221

Figure A – 94: DEPT-135 NMR spectrum of 133H+ in TFA-CDCl3

Figure A – 95: 125 MHz carbon-13 NMR spectrum of 133H+ in TFA-CDCl3

222

Figure A – 96: 500 MHz 1H NMR Spectrum of 117H+ in TFA-CDCl3

223

Figure A – 97: Selected nOe difference proton NMR spectra of 117H+ in TFA-CDCl3

10 9 8 7 6 5 4 3 2 1 ppm

224

Figure A – 98: 1H-1H COSY NMR spectrum of 117H+ in TFA-CDCl3

Figure A – 99: HSQC NMR spectrum of 117H+ in TFA-CDCl3

225

Figure A – 100: DEPT-135 NMR spectrum of 117H+ in TFA-CDCl3

Figure A – 101: 125 MHz carbon-13 NMR spectrum of 117H+ in TFA-CDCl3

226

Figure A – 102: 500 MHz 1H NMR Spectrum of 121H+ in TFA-CDCl3

227

Figure A – 103: 1H-1H COSY NMR spectrum of 121H+in TFA-CDCl3

Figure A – 104: HSQC NMR spectrum of 121H+ in TFA-CDCl3

228

Figure A – 105: DEPT-135 NMR spectrum of 121H+ in TFA-CDCl3

Figure A – 106: 125 MHz carbon-13 NMR spectrum of 121H+ in TFA-CDCl3

229

APPENDIX B

SELECTED UV-VIS SPECTRA

Figure B – 1: UV-Vis Spectra of 98 in dichloromethane (red line, free base), CH2Cl2 with 100

equiv TFA (green line, monocation), 50% TFA-CH2Cl2 (blue line, dication) and 5% DBU-

CH2Cl2 (purple line, monoanion)

0

2

4

6

8

10

12

14

16

300 400 500 600 700 800

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

230

Figure B – 2: UV-vis spectra of 120 in dichloromethane (red line, free base), CH2Cl2 with 5

equiv TFA (green line, monocation), 25% TFA-CH2Cl2 (blue line, dication) and 5% DBU-

CH2Cl2 (purple line, monoanion)

0

2

4

6

8

10

12

14

300 400 500 600 700 800

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

231

Figure B – 3: UV-vis spectra of 121 in 1% Et3N-CH2Cl2 (red line, free base) and in CH2Cl2

with 50 equiv. TFA (blue line, monocation)

0

1

2

3

4

5

6

7

300 400 500 600 700 800 900

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

232

Figure B – 4: UV-vis spectrum of 118 in chloroform

0

1

2

3

4

5

300 400 500 600 700 800

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

233

Figure B – 5: UV-vis spectra of 123 (red line) and 122 (blue line) in chloroform

0

1

2

3

4

5

6

300 400 500 600 700 800 900

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

234

Figure B – 6: UV-vis spectra of 117 in dichloromethane (red line, free base), 5% TFA-

dichloromethane (green line, monocation) and 5% DBU-dichloromethane (blue line)

0

5

10

15

20

25

300 400 500 600 700

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

235

Figure B – 7: UV-vis spectra of carbachlorin 133 in dichloromethane (red line), carbachlorin

monocation 133H+ in 5% TFA-dichloromethane and silver(III) chlorin 138 in dichloromethane

(purple line)

0

5

10

15

20

25

300 350 400 450 500 550 600 650 700

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

236

Figure B – 8: UV-vis spectra of 98 in dichloromethane with 0 equiv. TFA (red line), 2 equiv.

TFA (orange line), 5 equiv. TFA (green line), 10 equiv. TFA (light blue line), 20 equiv. TFA

(dark blue line) and 50 equiv. TFA (purple line)

0

1

2

3

4

5

6

7

8

9

10

300 400 500 600 700 800

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

237

Figure B – 9: UV-vis spectrum of 98H22+ in 50% TFA-CH2Cl2

Figure B – 10: UV-vis spectrum of deprotonated carbaporphyrin 98 in 5% DBU

0

2

4

6

8

10

12

14

16

300 400 500 600 700 800

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

0

1

2

3

4

5

6

7

300 400 500 600 700 800

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

238

Figure B – 11: UV-vis spectrum of 120 in dichloromethane

Figure B – 12: UV-vis spectrum of monoprotonated 120 in dichloromethane with 5 equiv TFA

0

0.5

1

1.5

2

2.5

3

3.5

4

4.5

5

300 400 500 600 700 800

Wavelength (nm)

εε εε x 1

0-4

M-1

cm

-1

0

1

2

3

4

5

6

7

300 400 500 600 700 800

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

239

Figure B – 13: UV-vis spectra of 120 in dichloromethane with 0 equiv. TFA (red line), 0.5

equiv. TFA (green line), 1 equiv. TFA (light blue line), 2 equiv. TFA (dark blue line) and 5

equiv. TFA (purple line)

0

1

2

3

4

5

6

300 400 500 600 700 800

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

240

Figure B – 14: UV-vis spectrum of diprotonated 120 in 25% TFA-CH2Cl2

Figure B – 15: UV-vis spectrum of 120 in 50% TFA-CH2Cl2

0

2

4

6

8

10

12

14

300 400 500 600 700 800

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

0

2

4

6

8

10

12

14

300 400 500 600 700 800

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

241

Figure B – 16: UV-vis spectrum of 120 in 1% Et3N-dichloromethane showing partial

deprotonation

Figure B – 17: UV-vis spectrum of monodeprotonated 120 in 5% DBU-CH2Cl2

0

0.5

1

1.5

2

2.5

3

3.5

4

4.5

5

300 400 500 600 700 800

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

0

0.5

1

1.5

2

2.5

3

3.5

4

4.5

5

300 400 500 600 700 800

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

242

Figure B – 18: UV-vis spectrum of 121 in 1% Et3N-CH2Cl2 (free base)

Figure B – 19: UV-vis spectrum of dimethylcarbaporphyrin 121 in 5% DBU-CH2Cl2

0

1

2

3

4

5

6

300 400 500 600 700 800 900

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

0

1

2

3

4

5

6

300 400 500 600 700 800 900

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

243

Figure B – 20: UV-vis spectrum of monoprotonated dimethylcarbaporphyrin 121 in in CH2Cl2

with 50 equiv. TFA

Figure B – 21: UV-vis spectrum of dimethylcarbaporphyrin 121 in 5% TFA-CH2Cl2

0

1

2

3

4

5

6

7

300 400 500 600 700 800 900

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

0

1

2

3

4

5

6

7

300 400 500 600 700 800 900

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

244

Figure B – 22: UV-vis spectrum of silver(III) carbaporphyrin 118 in chloroform

0

1

2

3

4

5

300 400 500 600 700 800

Wavelength (nm)

εε εε x 1

0-4

M-1

cm

-1

245

Figure B – 23: UV-vis spectrum of gold(III) carbaporphyrin 119 in chloroform

0

0.5

1

1.5

2

2.5

3

300 400 500 600 700

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

246

Figure B – 24: UV-Vis spectra on silver(III) carbaporphyrin (red line) and gold(III)

carbaporphyrin (blue line) in chloroform

0

1

2

3

4

5

300 400 500 600 700

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

247

Figure B – 25: UV-vis spectrum of nickel(II) carbaporphyrin 123 in chloroform

Figure B – 26: UV-vis spectra of carbachlorin by-product 117 in dichloromethane with 0 equiv.

TFA (red line), 0.5 equiv. TFA (green line), 1 equiv. TFA (blue line), and 2 equiv. TFA (purple

line)

0

1

2

3

4

5

6

300 400 500 600 700 800 900

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

0

5

10

15

20

25

300 400 500 600 700

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

248

Figure B – 27: UV-vis spectrum of carbachlorin by-product 133 in 5% DBU-dichloromethane

0

5

10

300 400 500 600 700

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

249

Figure B – 28: UV-vis spectrum of carbachlorin cation 133H+ in 1% TFA-dichloromethane

Figure B – 29: UV-vis spectrum of carbachlorin monocation 133H+ in 5% TFA-

dichloromethane

0

5

10

15

20

25

300 350 400 450 500 550 600 650 700

Wavelength (nm)

εε εε x 1

0-4

M-1

cm

-1

0

5

10

15

20

25

300 350 400 450 500 550 600 650 700

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1

250

Figure B – 30: UV-vis spectra of carbachlorin 133 in dichloromethane with 0 equiv. TFA (red

line), 0.5 equiv. TFA (green line), 1 equiv. TFA (blue line), and 2 equiv. TFA (purple line)

0

5

10

15

20

25

300 400 500 600 700

Wavelength (nm)

εε εεx 1

0-4

M-1

cm

-1